120
LATE HOLOCENE CLIMATE FLUCTUATIONS AT CASCADE LAKE, NORTHEASTERN AHKLUN MOUNTAINS, SOUTHWESTERN ALASKA By Kasey Kathan A Thesis Submitted in Partial Fulfillment of the Requirement for the Degree of Master of Science in Geology Northern Arizona University December 2006 Approved: ________________________________ Darrell S. Kaufman, Ph.D., Chair _________________________________ R. Scott Anderson, Ph.D. _________________________________ James Sample, Ph.D. _________________________________ Al Werner, Ph.D.

Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

Embed Size (px)

Citation preview

Page 1: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

LATE HOLOCENE CLIMATE FLUCTUATIONS AT CASCADE LAKE,

NORTHEASTERN AHKLUN MOUNTAINS, SOUTHWESTERN ALASKA

By Kasey Kathan

A Thesis

Submitted in Partial Fulfillment

of the Requirement for the Degree of

Master of Science

in Geology

Northern Arizona University

December 2006

Approved:

________________________________

Darrell S. Kaufman, Ph.D., Chair

_________________________________

R. Scott Anderson, Ph.D.

_________________________________

James Sample, Ph.D.

_________________________________

Al Werner, Ph.D.

Page 2: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

ii

ABSTRACT

LATE HOLOCENE CLIMATE FLUCTUATIONS AT CASCADE LAKE,

NORTHEASTERN AHKLUN MOUNTAINS, SOUTHWESTERN ALASKA

KASEY KATHAN

This study examines the relationships between glacier fluctuations, multiple

proxies from lake sediments and late Holocene climate fluctuations in southwestern

Alaska. Cascade Lake (3.1 km2; 60.1°N; 159.4°W; 330 m asl) is glacier-fed with 12% of

its drainage basin occupied by modern (1970) glaciers. The primary sediment contributor

is Cascade Glacier located 5.4 km up-valley. During the Little Ice Age, cirque and valley

glaciers of the Cascade Lake area experienced an average equilibrium-line altitude

lowering of 22 ± 12 m (n = 8) relative to 1957. Little Ice Age moraines stabilized ~1860

AD.

Three sediment cores were recovered from the lake and the lake was instrumented

with water-temperature loggers and sediment traps from 2004-2006. 137Cs, 210Pb profiles,

the Aniakchak tephra (3500 cal yr BP) and the regional tephra (Tephra B; 6100 cal yr

BP) provide limited age control and indicate an average late Holocene sedimentation rate

of ~0.07 cm/yr. The cores are composed of rhythmically laminated mud, with an average

lamination thickness of 0.4 cm; these laminations do not represent annual sedimentation

(e.g. varves).

Biogenic silica was analyzed at an average spacing of 0.2 cm (~3 yr) over the last

150 yr and is well correlated with spring temperature (r = 0.70, p = 9.16 x 10-5),

precipitation (Nuyakuk discharge; r = 0.73, p = 2.2 x 10-6) and the Aleutian Low Pacific

Page 3: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

iii

Index (r = 0.68, p = 1.37 x 10-7). These relationships suggest a link to nutrient input to

the lake. High BSi flux can generally been related to mild conditions and low BSi flux to

cool conditions. This interpretation and other proxy data indicate mild climates at ~7000

cal yr BP (Holocene thermal maximum) and 1600-1400 cal yr BP (Medieval Warm

Period?). Cool conditions persisted between ~3000-2500 cal yr BP (onset of

Neoglaciation) and 1400 cal yr BP to 1870 AD (Little Ice Age). The pattern of climate

variability in Cascade Lake generally agrees with widely recognized Holocene climate

fluctuations, though the ages differ. This may be due to a time-transgressive nature of

these climate events or, more likely, a poorly constrained age model.

Based on the interpretation of several periods of climate variability in the Ahklun

Mountains during the late Holocene, the climate of the 20th century can be placed in

context. Within the Cascade Lake record, no other period has comparable biogenic

silica) concentrations to the 20th century. This suggests an unprecedented strengthening

of the Aleutian low during this most recent period.

Page 4: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

iv

ACKNOWLEDGMENTS

It was through the hard-work and help of many people that this project was

completed. My sincere gratitude first goes to Darrell Kaufman, as he has been a

remarkable advisor and role-model throughout every step of this project. Without his

energy during field-work, excitement over new data and encouragement during the

writing process, this thesis would lack the integrity and detail that it now has. I thank

him for always having an open office door, even when his schedule was busy, and for

asking the questions that kept me motivated.

My thanks also extend to my committee members, Jim Sample, Scott Anderson,

and Al Werner. Their comments and observations helped to tie my loose ends together.

I’ve always enjoyed our conversations both within the formal committee meeting setting

and informally in the hallway. I appreciate the extensive time and brainpower that you

poured into my thesis. I especially appreciate the help of Al Werner, because it was

through his encouragement that I choose to continue my education beyond Mount

Holyoke College.

There are many additional people who have contributed to this project with their

time, and expertise. Tom Daigle, Erin Dopfel, Mike Hamilton, Heidi Roop, Apayo

Moore, Al Werner and of course Darrell Kaufman provided much needed assistance and

ideas in the field, while Togiak National Wildlife Refuge provided the logistical support.

Feng Sheng Hu of University of Illinois and Richard Ku of University of Southern

California provided the 210Pb and 137Cs analyses and John Southon of University of

California, Irvine for the radiocarbon analysis. Jim Wittke spent a significant amount of

Page 5: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

v

time helping analyze tephra on the electron microprobe and scanning electron

microscope. Tom Daigle, Jordan Bright, Caleb Schiff, Nick McKay and Chris Kassel

each get thanked for spending time talking over ideas with me and asking the hard

questions.

Special thanks go to my family and friends for putting up with me. Sheena

Styger, Chloe Bonamici, Christyanne Melendez, and Monica Hansen provided necessary

distractions and coffee breaks. My parents have always supported their wandering child

and continue to be the most significant role models in my life. Skip Martin has provided

me with retreats to cabins in the woods and many hugs from 3000 miles away.

This project was funded by National Science Foundation grants ARC-0455043

and ATM-0318341 awarded to Dr. Darrell Kaufman, and a Geological Society of

America graduate student research grant.

Page 6: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

vi

TABLE OF CONTENTS

Page ABSTRACT………………………………………………………………………………ii ACKNOWLEDGMENTS………………………………………………………………..iv LIST OF TABLES………………………………………………………………...……..vii LIST OF FIGURES……………………………………………..………………………viii CHAPTERS

1. INTRODUCTION………………………………………………………………1 a. Project Objective 1 b. Background 2 c. Study Area 8

2. MODERN CLIMATE AND SEDIMENTATION AT CASCADE LAKE………9

a. Methods 9 b. Results 11 c. Discussion 13 d. Summary 15

3. GLACIAL GEOLOGY…………………………………………………………16 a. Background 16 b. Methods 18 c. Results 20 d. Discussion 23 e. Summary 26

4. CASCADE LAKE SEDIMENTS………………………………………………..26 a. Background 26 b. Methods 27 c. Results and Discussion 32

5. BIOGENIC SILICA AND CLIMATE INTERPRETATIONS…………...……..45 a. Methods 47 b. Results 48 c. Discussion 49

6. CONCLUSIONS………………………………………………………………...55

Page 7: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

vii

LIST OF TABLES

Table Page

1. Instruments deployed at Cascade Lake………………………………….68

2. Data from sediment traps………………………………………………..69

3. Climate comparison with lichen growth-curve sites…………………….70

4. Lichen sizes and ages……………………………………………………71

5. Equilibrium line altitudes for glaciers…………………………………...72

6. Lake cores recovered for this study……………………………………...73

7. Tephra descriptions from Core 02……………………………………….74

8. BSi and meteorological correlations……………………………………..75

Page 8: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

viii

LIST OF FIGURES

Page

1. Summary of Holocene climate fluctuations………………………………... 76

2. Map of Alaska and the Ahklun Mountains………………………………… 77

3. Color-infrared aerial photograph of the study area……………………….... 78

4. Types of instruments deployed for this study……………………………… 79

5. Data recovered from 2004-2005, 2005-2006 sediment traps………………. 80

6. Air and water temperature data from the study area……………………….. 81

7. Relationship between Dillingham, Bethel and Cascade temperatures…….. 82

8. Cascade glaciers…………………………………………………………… 83

9. Equilibrium-line-altitude curves…………………………………………… 84

10. Moraines of Cascade Glacier and the East Inflow cirque………………….. 85

11. Lichen growth curves………………………………………………………. 86

12. Retreat rates of Cascade Glacier……………………………………………. 87

13. Bathymetric map…………………………………………………………… 88

14. Methods used in lamination counting……………………………………… 89

15. Acoustic profile from Core 02 vicinity…………………………………….. 90

16. Lithostratigraphic logs…………………………………………………….. 91

17. Correlations with long and surface………………………………………… 92

18. Test pit wall………………………………………………………………… 93

19. 210Pb and 137Cs profiles…………………………………………………….. 94

20. Geochemical data from the 2.85 m tephra and Aniakchak reference samples. 95

21. Scanning Electron Microscope images of tephra shards………………….. 96

22. Age model…………………………………………………………………. 97

23. Magnetic susceptibility…………………………………………………….. 98

24. Organic matter content and bulk density…………………………………... 99

25. Lamination thickness……………………………………………………… 100

26. Biogenic silica…………………………………………………………….. 101

27. Biogenic silica of the last 2200 years……………………………………… 102

28. Core 02 proxy data by age…………………………………………………. 103

Page 9: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

1

CHAPTER 1:

INTRODUCTION

The high latitudes have been shown to be particularly sensitive to global climate

change, with dramatic environmental changes induced by subtle climate variations

(Overpeck et al., 1997). Throughout the Quaternary, inter-annual and century-scale

climate variation has been the norm; however, the timing, synchronicity and forcing

mechanisms of these fluctuations are still poorly understood and necessitate further work.

The investigation and interpretation of these past environmental changes provide a

context within which to place our understanding of present and future climate

fluctuations.

Lake sediment cores have proven particularly successful in providing valuable

interpretations of Holocene climate fluctuations (Pienitz et al., 2004). Lake sediments

provide continuous records of glacial fluctuations, lake productivity, sediment fluxes and

a wide variety of other environmental changes induced by climate variability. Using

proxy indicators available in lacustrine sediments, it is possible to determine the range of

non-anthropogenic Holocene climate variability. The naturally induced events of the

Holocene (e.g., Neoglaciation, Medieval Warm Period, and Little Ice Age) can then be

compared with the twentieth-century climate changes that are related to anthropogenic

forcing.

Project Objective

The overarching goal of this study is to determine the range of climate variability

preserved in the sediments of Cascade Lake, Alaska. This is accomplished by: (1)

Page 10: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

2

establishing an understanding of the modern and historic climate and sedimentation

patterns at Cascade Lake, (2) determining the recent glacial fluctuations within the

drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and

fluctuations in Cascade Lake sediment cores, and (4) using proxy indicators to interpret

Holocene climate variations.

Background

Holocene Climate Variability. Initial paleoclimate reconstructions from the marine and

ice cores suggested that the Holocene was a period of relative stability compared to the

large-scale climate fluctuations of the Last Glacial Maximum. Modern alpine glaciers

were originally assumed to be merely the remnants of maximum Pleistocene ice extent.

However, it is now recognized that this ‘quiescent’ period includes a series of

pronounced climate fluctuations (e.g., Mayewski et al., 2004). Multiple cooling events

have been identified that punctuate the current interglacial period; perhaps the most well

known is the 400-year-long Little Ice Age (LIA). This period of multiple “glacial

expansion(s) subsequent to the maximum Hypsithermal shrinkage” has been termed

Neoglaciation (Porter and Denton, 1967).

One of the first studies to systematically approach North American climate

fluctuations of the late Holocene was by Denton and Karlén (1973). Their study of the

expansion and contraction of 51 North American and European alpine glaciers

recognized three prominent intervals of glacial expansion: 3300-2400, 1250-1050 cal yr

BP (where “BP” = before 1950 AD), and the Little Ice Age (1500-1900 AD). This

millennial-scale pattern (~1500 and 2000-2800 year climate periodicity) of widespread

Page 11: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

3

late Holocene cooling events is gaining support as evidence of similar climate variability

emerges from other high-resolution records (Bond and Showers, 1997; Campbell et al.,

1998; Bianchi and McCave, 1999; Leonard and Reasoner, 1999; Noren et al., 2002;

Mayewski et al., 2004). Mayewski et al. (2004) summarized this supporting evidence in

a study of ~50 spatially distributed paleoclimate records. They determined that a

minimum of six global Holocene cooling events are associated with this pattern occurring

at 9000-8000, 6000-5000, 4200-3800, 3500-2500, 1200-1000 and 600-150 cal yr BP

(Figure 1).

The millennial-scale climate variability of the Holocene is not only marked by

cool periods. Mild, warm and dry periods have also been recognized at 10600-9300,

7900-6300, 2700-1500, 960-610 cal yr BP (Figure 1, O’Brien et al., 1995). The warm

period of approximately 1000 cal yr BP (1000-1300 AD) has been termed the Medieval

Warm Period (MWP). This period is often cited as being similar to the modern 20th

century warming; however, a compilation by Crowley and Lowrey (2000) determined

that the MWP was only ~0.20°C warmer than the LIA. The spatial and temporal patterns

and the amplitude of this warm period must be resolved before it can be used as a model

for 20th century warming.

A forcing mechanism for this millennial-scale climate periodicity has yet to be

determined. Some evidence suggests that these cycles are related to Dansgaard-Oeschger

like ice-rafting events, and therefore linked to North Atlantic thermohaline circulation

and a strongly coupled ocean-atmosphere system (Alley et al., 1997; Bond and Showers,

1997). Other evidence suggests that external factors, such as persistent solar variability,

Page 12: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

4

explain the widespread distribution of these climate events (Denton and Karlén, 1973;

Mayewski et al., 1997; Bond et al., 2001).

Neoglaciation in Alaska. The glacial records of past climate variability in Alaska

generally concur with the more widely recognized events of the late Holocene. Calkin’s

(1988) summary of Holocene glaciation in Alaska focused on glacial moraines that

indicate four major periods of glacial advance including: an initial advance between

6000-5000 cal yr BP, two distinct advances at 3600-2000, and 1700-1100 cal yr BP and

the most recent advance of the Little Ice Age at 750-400 cal yr BP. While these periods

of glacial advance have been recognized throughout Alaska, their timing, duration and

magnitude vary considerably. For instance, the maximum Holocene ice extent in regions

of the Kenai Peninsula was achieved during the 3600-2000 cal yr BP advance whereas

many other local areas experienced maximum ice extent during the LIA (Calkin, 1988).

Some of the variation in the timing of Neoglacial advances could be attributed to

the difficulty in identifying geomorphic evidence for the extent of early Neoglacial

advances, which might have been obliterated by more extensive LIA advances.

However, supporting evidence has been derived from proxy records such as pollen, lake-

level and tree-ring studies. In southern Alaska these supporting studies indicate that

periods of glacial advance identified in glacial geomorphic studies correspond to periods

of increased wetness (e.g., Hansen and Engstrom, 1996; Abbott et al., 2000; Calkin et al.,

2001).

Pacific Ocean and Alaska Climate. Interdecadal climate variability in the North Pacific

region has been well documented throughout the last century and likely influences the

Page 13: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

5

millennial-scale climate fluctuations of the Holocene ( Minobe and Mantua, 1999;

Schneider and Cornuelle, 2006). Although the mechanism for this interdecadal climate

variability is unknown, a basin-wide phenomenon that originates in the tropics and is

closely related to the shorter-term oscillations of sea-surface temperature and sea-level

pressure has been suggested (D’Arrigo et al., 2005). The primary two short-term

oscillations are the El Niño-Southern Oscillation (ENSO) and the Aleutian low-pressure

cell (AL). ENSO regimes typically persist for 6-18 months and are responsible for shifts

in the locations of warm tropical waters. These shifts in sea-surface temperatures (SST)

in turn affect the regional precipitation and mean annual temperatures (Papineau, 2001;

Neal et al., 2002). The dominant mode of North Pacific wintertime atmospheric

circulation is the Aleutian low. A deepened AL results in a strengthened southerly flow

over the eastern Pacific, and warmer temperatures along the western coast of North

America (Rodionov et al., 2005). Each of these components contributes to the Pacific

Decadal Oscillation (PDO; measured by SST), which is one of the best summaries of

North Pacific interdecadal climate variability.

The PDO has its primary signature in the North Pacific Ocean with a secondary

imprint in the tropics (where the ENSO fingerprint is primary) and a periodicity of

approximately every 20-30 yr. Over the historic period the PDO has had several

documented shifts in its primary modes (Kayano et al., 2005). ‘Cool’ phases (cold waters

in the Gulf of Alaska) persisted from 1900-1924 and 1947-1976. Subsequent ‘warm’

phases occurred from 1925-1946 and 1977 to the mid-1990s. These interdecadal shifts in

the PDO have a strong link to interannual variation of the AL. Warm modes of the PDO

are associated with a general strengthening of the AL; however, the interannual variance

Page 14: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

6

of the AL during the warm PDO phases is two to three times greater than other periods

(Minobe and Mantua, 1999). Though these patterns of variability in the closely coupled

ocean-atmosphere system of the North Pacific have only been documented over the last

100 yr, paleoclimate studies suggest that these patterns have persistently influenced

southern Alaska climate throughout the Holocene. Generally, the association between

increased wetness and periods of glacial advances in northwestern North America implies

a strong influence of the AL on Holocene glaciation (Spooner et al., 2003; Anderson et

al., 2005).

Previous Work in Ahklun Mountains, Alaska. Although significant paleoclimate research

has been completed in Alaska, relatively little attention has focused on the late Holocene

in the southwestern portion of the state. Previous work in southwestern Alaska has

largely focused on the Ahklun Mountains. The Ahklun Mountains are a 250-km-long

mountain belt trending southwest to northeast between Bristol Bay and Kuskokwim Bay

lowlands (Figure 2). They form the largest mountain range west of the Alaska Range and

harbored a large ice cap during the late Wisconsin that was separate from the Cordilleran

Ice Sheet. Most of the paleoclimate research completed in the region has focused on

Pleistocene and Younger Dryas glacial advances (Briner and Kaufman, 2000; Kaufman et

al., 2001; Manley et al., 2001; Briner et al., 2002; Hu and Shemesh, 2003; Hu et al.,

2006).

Studies of the Holocene climate in the Ahklun Mountains remain rather limited.

A lacustrine sediment sequence recovered from Arolik Lake in the southwestern Ahklun

Mountains produced a 33,000 yr record of climate variability (Kaufman et al., 2003).

While high climate variability was observed in the Arolik Lake record, sampling

Page 15: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

7

resolution was too coarse to resolve Holocene fluctuations. A more detailed study (Hu et

al., 2003) observed regular oscillations in lake productivity coinciding with solar

variability, but this study focused on a portion of the core ranging in age from 12,000 to

2300 cal yr BP and did not address Neoglacial cooling events.

Studies that directly address late Holocene paleoclimate in the Ahklun Mountains

include a pollen study by Axford and Kaufman (2004) and a combined lacustrine/glacial-

moraine study by Levy et al. (2004). While Little Swift Lake, in the northwestern

Ahklun Mountains, does not contain clear evidence of specific Neoglacial cooling events,

the pollen record does indicate progressive climatic deterioration (cooling) throughout the

late Holocene (Axford and Kaufman, 2004). Levy et al.’s (2004) work on Waskey Lake

is the most complete work to date on Ahklun Mountain Neoglaciation. They identified

multiple late Holocene glacial advances by the presence of several discrete end moraines

in the forefields of the modern glaciers. Both the geomorphic evidence and the lake-

sediment record indicate that the peak Neoglacial activity occurred at Waskey Lake

between 700 and 200 cal yr BP. The lake sediments indicate an onset of Neoglaciation at

3100 cal yr BP. This age is significantly younger than the ages suggested for major

Neoglacial advances across the rest of Alaska (Brooks Range, Kenai Mountains, and

Seward Peninsula). This suggests that the Ahklun Mountains may be slightly out of

phase with widespread climate events of the late Holocene. Without additional records of

Neoglaciation in the Ahklun Mountains, however, it is difficult to determine the full

significance of the Waskey Lake record. If the timing of glacial activity at Waskey Lake

matches other places across the Ahklun Mountains this would imply that an unknown

regional-scale climate forcing exerts a stronger influence than the large-widespread

Page 16: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

8

climate forcing affecting other areas of Alaska and potentially the globe. Cascade Lake is

located only 38 km from Waskey Lake providing an opportunity to test the

reproducibility of the Waskey Lake Holocene climate record reported by Levy et al.

(2004).

Study Area

Presently, over 100 active glaciers are located in the northern portions of the

Ahklun Mountains where numerous peaks are over 1500 m a.s.l. The largest of these

northern glacial complexes are Chikuminuk Glacier (~5.6 km2) and an unnamed glacier

(hereafter, “Cascade Glacier”, ~3.4 km2) to the west and south respectively (Figure 3).

Cascade Lake (60° 8’ 11.9’’N; 159° 23’ 57.3’’W) lies 5.4 km downstream from the

primary outflow of Cascade Glacier at 335 m elevation.

The drainage basin area of Cascade Lake is 43.7 km2, 12% of which is occupied

by active glaciers. The area is vegetated with stands of alder and willow shrub and upland

tundra. Cascade Lake (3.1 km2) has multiple inflows, both melt-water streams and non-

glacial channels, though the primary sediment contributor is Cascade Glacier (Figure 3).

Cascade Lake is formed in late Wisconsin drift underlain by Jurassic volcaniclastic

sandstones, conglomerates and argillites that are members of the Gemuk group (Box et

al., 1993). Two interpreted faults bisect the field area, one crossing the northern portion

of the lake and the other crossing mid-way down the melt-water channel of Cascade

Glacier (Hoare and Coonrad, 1959).

Page 17: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

9

CHAPTER 2: MODERN CLIMATE AND SEDIMENTATION AT CASCADE LAKE

In order to make quantitative interpretations of Holocene climate variations

preserved in lake sediments, it is necessary to understand the climate factors affecting

sediment transfer and deposition. Factors such as temperature, rainfall, and snowmelt

intensity, have a direct impact on the sediment characteristics preserved in lakes (Leeman

and Niessen, 1994; Hardy et al., 1996; Noon et al., 2001; Francus et al., 2002). A small-

scale study of the modern sedimentation and climate at Cascade Lake was undertaken to

enable a better understanding of the influence of climate on its sediment record and to

evaluate the potential for annual sedimentation patterns.

Methods

Moorings. Instrument moorings were deployed at water depths of 24.2 m (mooring 1)

and 52.4 m (mooring 2) at glacial distal and proximal sites, respectively, within Cascade

Lake (Figure 3, Table 1). Each mooring consisted of a rock anchor, two sediment traps,

two temperature loggers and a buoy. The sediment traps were made from 12.5-cm-

diameter funnels that fed rigid lexan tubes with an inner diameter of 1.6 cm (focusing

factor = 62 times) (Figure 4). Traps were placed at ~1 m above the lake bottom and ~3 m

below the lake surface. Buoys were set at ~2 m below the lake surface to avoid exposure

during lower winter lake levels or freezing in winter ice. Each sediment trap was paired

with an Onset Water Temp Pro temperature logger that recorded temperature every hour

beginning June 28, 2004.

Page 18: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

10

Trapped Sediments. The sediment traps were recovered on August 10 and 11, 2005 and

on July 19th, 2006. The sediment was described and sub-sampled into stratigraphically

distinct units visible in the receiving tube. The sediment from the 2005 traps at mooring

2 was analyzed for bulk-sediment density, and 2005 sediment from mooring 1 was

processed for organic-matter content (OM), grain size, and biogenic silica content (BSi).

The 2006 sediment from mooring 2 was analyzed for bulk-sediment density and BSi

content.

Sediment grain-size distribution was determined using 0.25 cm3 samples analyzed

by a Coulter LS230 laser diffraction analyzer. Samples were pre-treated with 30% H2O2

to remove organics and 1M NaOH to dissolve BSi, and stored in sodium

hexametaphosphate (a dispersing agent) to prevent flocculation prior to analysis. Each

sample was analyzed four times for 80 seconds and all runs were averaged for the

samples final grain-size estimate. OM content was determined by percent loss-on-

ignition (LOI) and BSi analysis was completed using the methods of Mortlock and

Froelich (1989, see Lake Sediments Methods).

Air Temperature. Onset Hobo temperature loggers were installed on the western shore of

Cascade Lake and ~0.6 km from the modern Cascade Glacier front (Figure 3). The

lakeshore logger was placed inside a solar radiation shield and mounted at a height of ~3

m on a large willow tree in a stand of alder (Figure 4). The logger recorded air

temperature every hour for a period of June 28, 2004 to August 9, 2005, and from August

12, 2005 to July 20, 2006. The glacier front air temperature logger was mounted on

stabilized copper poles ~2 m above the ground and recorded temperature every hour from

August 13, 2005 to July 21, 2006.

Page 19: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

11

Results and Discussion

Sediment Traps. Mooring 1 was recovered from its deployed location in both 2005 and

2006; mooring 2 was recovered from deeper water than which it was originally deployed

in 2005. The mooring likely moved (drifted or was carried by lake ice) during the

collection year, but the sediment traps and temperature loggers were all intact and

seemingly undisturbed. Upon recovery the top traps were covered in a thick algal mat

and had collected limited porous sediments with live organisms. The bottom traps

contained significantly more sediment than the top traps (2005 mooring 2 had slightly

overflowed into the collection tube) and were sub-sampled into an average of five

stratigraphic units (Figure 5). Sediment in the bottom traps was typically gray to gray-

green with organic material overlaying dense mud with gradational contacts between

units.

Significantly less sediment accumulated during 2005-2006 than in 2004-2005.

Unfortunately, a fingerling fish was caught in the receiving tube of 2006 mooring 1 and

made field collection of the sediment impossible. However, mooring 2 collected 8.5 cm

of sediment in 2006 compared to the 18.1 cm in 2005. The average sediment

accumulation in the bottom trap of mooring 2 was 0.10 g/cm2/yr in 2004-2005, and 0.03

g/cm2/yr in 2005-2006, a 70% reduction (Tables 1 and 2). Because of possible

complications caused by the movement of mooring 2 during the year and the overflowed

collection tube, I focused on 2004-2005 mooring 1 for detailed analysis. Each of the

stratigraphic units was analyzed for OM, grain size, and BSi content. The 2006 mooring

2 sediments were also analyzed for bulk density and BSi content for comparison (Figure

Page 20: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

12

5). OM, grain size, and BSi content all covaried in the sediment from mooring 1 in 2005,

with their lowest values in the middle units (Figure 5). The sediment from mooring 2 in

2006 contained less BSi overall (16% reduction in BSi percentage) and did not show

decreased BSi content in the center units. Adjusting BSi values for the dilution caused by

the increased sedimentation in the 2004-2005 collection year, results in a 77% reduction

in BSi flux during 2005-2006 (Table 2).

The results from the bottom trap of 2004 mooring 1 exhibit a seasonal pattern

(Figure 5). Because the traps were not equipped to record the timing of sediment

deposition within the collection tubes, the seasonality of each lithologic unit is not

known. Nonetheless, the decrease in OM, BSi, and grain size in the central part of the

receiving tube of mooring 1 is indicative of winter sediment deposition in a glacial-fed

lake. The increase in BSi in the uppermost sediments (units 1 and 2) and the coarse

sediments (coarse silt) in unit 1 might indicate higher stream discharge associated with

spring melt-out. Mooring 1 was located farther from the glacial melt-water inflows, and

therefore was less sensitive to the glacial melt-water influences and may not have

captured the seasonal signal as well as the more proximal site might have.

The 70% decrease in sediment flux and the 77% reduction in BSi deposition

between the two collection years are indicative of a strong influence of interannual

weather differences on sediment transport and accumulation. Spring and summer (May

to September) surface-water temperatures (air temperature not available for 2005) were

~1.5 °C colder in 2006 than in 2005, and in July of 2006, significantly more snow

remained on the landscape. Additionally, less (no quantitative measurements exist) melt-

water occupied the channels of the Cascade Glacier outwash plain than in the previous

Page 21: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

13

year. There is no precipitation data for Dillingham post-1998; however, precipitation

data from Bethel over the collection years shows a 7.6 cm decrease in annual

precipitation from 2005 to 2006 (62.5 vs. 54.9 cm; WRCC, 2006). Possibly the main

pulse of sediment associated with peak melt-water discharge had not occurred by the time

the traps were recovered in July 2006. More likely, the cooler spring/early summer

temperatures and decreased precipitation retarded melt-water production (i.e. peak

discharge) and decreased the sediment flux and associated nutrients for algal growth.

Water Temperature. The four temperature loggers were recovered in August 2005 and

July 2006 (Figure 6). The water temperature data indicate that Cascade Lake is dimictic

with fall mixing occurring in late October and spring melt-out and mixing occurring in

late May to early June. Thermal mixing during the fall lasted 21 days in 2004 and 13 days

in 2005 whereas spring mixing took place over 8 days in 2005 and 4 days in 2006. Fall

cooling and spring warming leading to or following overturning were gradual at a rate of

about 1 °C per 9 days. The lake-surface temperatures are similar at both mooring sites

(0.1 °C warmer at the mooring 1 site). During the ice-free months, the lake-surface

temperatures track air temperature with a minimal lag time (hours) and average 5 ± 9 °C

cooler (based on hourly averages).

Air Temperature. In 2005, the lakeshore temperature logger malfunctioned and stopped

recording on December 22, 2004 (after 176 days, Figure 6). A new lakeshore logger and

the glacier terminus logger were deployed in August 2005 and downloaded in July 2006.

The radiation shield at the glacier terminus was knocked off its mount and the logger

Page 22: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

14

recorded a stable temperature of ~-7 °C from the end of January to the beginning of June

2006, probably because the fallen logger was buried by snow during this interval.

Freezing air temperatures at the lake were first reached in September and

generally remained below freezing by November in both 2004-2005 and 2005-2006.

Temperatures averaged above freezing starting in early May. For the period successfully

recorded by both lakeshore air temperature loggers (August 12-January 22), the average

2004 temperature was 1.2 °C and the average 2005 temperature was –1.4 °C. The coldest

recorded temperature in 2004 was –28 °C in early December, and in 2006 it was –40 °C

in late January. The onset of freezing air temperatures and the timing of above-freezing

temperatures in the spring correspond well to the build-up and melt-out of lake ice, just

prior to and after lake overturning. Air temperatures lead lake ice-up and melt-out by

about a month.

The glacier logger first recorded freezing temperatures two months earlier than

the lake, and minimum temperatures occurred just before it went isothermal (-35°C,

Figure 6). During the summer (JJA) the lakeshore air temperature averaged 2 ± 1 °C

warmer than the glacier, whereas the lakeshore was 2 ± 4 °C cooler during the winter

(DJF). With an elevation difference of 340 m these two temperature stations record an

environmental lapse rate of ~6 °C/km during the summer and an inverted profile in the

winter.

The air temperature logged at Cascade Lake can be compared with other climate

stations in the region. The goal of this comparison was to determine whether the longer-

term temperature records of the climate stations could confidently be extrapolated to

Cascade Lake. The two closest climate stations are Bethel (~148 km to the northwest)

Page 23: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

15

and Dillingham (~140 km to the southeast), where daily maximum and minimum

temperature data are available (PNWPEST, 2006). I compared the daily maximum,

minimum, and average daily temperatures from the climate stations (mean of two values)

and from the Cascade Lake hourly air-temperature logger (mean of 24 values) and found

the strongest correlation between the average daily temperature (r = 0.97 and 0.96 for

Dillingham and Bethel, respectively). Daily temperature at Cascade Lake averaged 3.5 ±

3.1 °C (n = 498) cooler than Dillingham and 1.1 ± 3.4 °C cooler than Bethel (Figure 7).

The temperature differences between Cascade Lake shoreline and Dillingham vary by

season. During the summer, Cascade Lake was 2 °C cooler than Dillingham and 5 °C

colder in the winter (Figure 7). The temperature differences between Dillingham and

Cascade Lake indicate a summer environmental lapse rate of 7 °C/km. This is higher

than the typical lapse rates of 6 °C/km (Porter, 2001), but similar to that of the lake and

glacier temperature loggers. The somewhat higher lapse rate may be due to the more

southerly and coastal location of Dillingham. Although temperature at both Dillingham

and Bethel are strongly correlated with Cascade Lake, I chose to focus on the climate

record at Dillingham because the temperature record there extends back to 1919,

(compared with 1949 at Bethel).

Page 24: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

16

CHAPTER 3: GLACIAL GEOLOGY

Glaciers are sensitive to climate variations and are known to have fluctuated

during the late Holocene in mountains across Alaska (Calkin, 1988; Calkin et al., 2001).

The study of glacial deposits can provide information on the timing and magnitude of

climate change that caused their deposition. By combining a glacial-geologic and a lake-

sediment study from the same drainage, the two complimentary records of climate

variability can be compared.

Background

Lichenometry. Lichenometry is a useful dating method applied to late Holocene

landforms (Locke et al., 1979). In many cases, the historic record is too fragmented,

carbonaceous material is absent or the landforms are too young (<500 yr) for reliable

radiocarbon dating, or the tree-ring record is too short. Lichenometry is based on the

premise that lichens grow at a predictable rate once they colonize a surface. Rhizocarpon

geographicum and R. alpicola are the two most commonly used species because they

have slow growth rates (0.02-2.0 mm/yr; Innes, 1985). These two species have slightly

different growth rates and they are impossible to distinguish in the field (Werner, 1990).

Most authors report data as Rhizocarpon geographicum senso lato (s.l.) and assume that

this includes measurements of both R. geographicum and R. alpicola thalli.

The growth rate of Rhizocarpon is dependent on a variety of factors including, but

not limited to: substrate lithology, moisture availability, temperature, altitude, and

Page 25: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

17

snowcover (Innes, 1985). Each of these factors can vary significantly from one location

to another, therefore, it is important to use a growth curve developed either from the

region of interest, or from a similar environment. Lichen ages reflect the timing of

landform stabilization and because moraines take some time to stabilize following the

retreat of abutting glacial ice, they provide minimum ages for moraine deposition.

Additionally, actively eroding landforms can not produce reliable lichenometric ages and

should not be used (O’Neal, 2006).

Equilibrium-Line Altitude. The equilibrium-line altitude (ELA) serves as one of the most

useful parameters to quantify the effect of climate on a glacier. The ELA separates the

areas of accumulation and ablation and is controlled by a variety of factors including

glacier hypsometry, glacier aspect, and other local climate conditions (Torsnes et al.,

1993). Changes in ELA over time are generally most sensitive to winter precipitation

and summer temperature, thereby enabling effective paleoclimate reconstructions (Dahl

and Nesje, 1992).

A variety of methods have been developed to reconstruct former ELAs. Torsnes

et al. (1993) compared various methodologies and determined that the accumulation area

ratio (AAR) method resulted in the least error when calculating the ELA of modern

glaciers with ELAs known through mass-balance measurements. The accumulation area

ratio is defined as the ratio of the accumulation area to the total glacier area and is used to

determine steady-state ELA. Meier and Post (1962) demonstrated that the AAR for

steady-state glaciers in northwestern North America ranged from 0.5 to 0.8. The average

of these values (0.65) is often used in ELA reconstructions (Porter, 2001).

Page 26: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

18

Methods

Lichenometry. Glacial and fluvial deposits believed to be Holocene in age were identified

on aerial photos and then studied in the field, or in some cases deposits were first

recognized in the field. These deposits included moraines of the most accessible glaciers

in the study area, fluvial terraces in the outwash plains of these glaciers, and stable

surfaces that were formed during glacier retreat. The long and short axes of large

Rhizocarpon geographicum s.l. thalli were measured using digital calipers. Landform

surfaces were searched until no larger thalli could be found. This involved searching the

entire landform surface and measuring approximately 15 lichen thalli. Generally, the

largest lichens were found on stable slopes (crests of moraines) and in protected

depressions on landform surfaces. Lichens with a long axis that measured over 20%

larger than the short axis were rejected because of this non-regular growth pattern (Calkin

and Ellis, 1980).

Because a lichen growth curve has not been developed for the Ahklun Mountains,

the lichen sizes were compared to the growth rates at five other Alaska locales including:

the Brooks Range, Seward Peninsula, central Alaska Range, Wrangell-St. Elias

Mountains, and southern Kenai Mountains (Figure 2). Solomina and Calkin (2003)

recently revised the growth curves for each of these areas based on additional control

points, newly calibrated 14C ages, and standardized fitting routines. These growth curves

either used the single largest thallus diameter or the averages of the five largest

diameters. Both of these values were calculated for the Cascade River valley data and

compared with the corresponding growth curves. Uncertainty in the derived ages results

Page 27: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

19

from the variety of climates represented by the five growth rates and their dissimilarities

from the Cascade River valley and measurement errors. Of the five areas for which

Solomina and Calkin (2003) developed growth curves, the mean annual and summer

temperature (and likely the precipitation) of the Cascade Lake area are most similar to

Seward Peninsula and therefore was used for age assignments (Table 3).

Equilibrium-Line Altitudes. Eight glaciers were chosen for ELA analysis, including those

visited during the 2005 field season, and those associated with laterally extensive and

prominent moraines on the 1957 aerial photographs. These eight glaciers include the

Cascade and Chikuminuk valley glaciers and six cirque glaciers (Hungry Joe Cirque,

Divided Cirque, East Inflow Cirque, Cirque 1, Cirque 2, Cirque 3, Figure 8). Four of the

glaciers are within the Cascade Lake drainage basin and all are located within 9 km of the

lake.

I used the AAR method for ELA reconstructions with an AAR value of 0.6 to

facilitate direct comparison with the Waskey Lake area (Levy et al., 2004). The extent of

ice in the 1957 aerial photographs and the reconstructed LIA ice limits (based on field

glacial-geomorphic evidence and aerial photograph interpretation) were plotted on a

1:63,360-scale topographic map (1973) in ArcGIS v. 9 (i.e., ‘modern’ = 1957). This

assumes that the 1973 contours were a reasonably accurate representation of the 1957

glacier surface. The area of each glacier was measured at ~100-ft-altitude intervals for

both the modern and reconstructed LIA ice limits. A multi-order polynomial was fit to

the data and used to determine the altitude corresponding to an AAR of 0.6 (40% of the

total area lies within the zone of ablation; Figure 9).

Page 28: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

20

Results and Discussion

Lichenometry. A total of 160 lichens with circular thalli were measured on 13 deposits in

the forefield of five extant glaciers (Figure 8, Table 4). The outer moraines represent the

largest Neoglacial advance while other deposits further up-valley represent smaller

subsequent readvances, delays in retreat, or are deposits of a non-glacial origin (debris-

flow deposits, river terraces, etc.). Of the five forefields, only two contained stable late

Holocene moraines with measurable lichens, the other deposits were either unstable

(slumped) over this period, or could not be clearly associated with the maximum ice

advance . Lichens were measured on two distinct moraine deposits on both the left and

right lateral sides of Cascade River valley and the terminal moraine of the East Inflow

cirque (Figures 8, 10). The age estimates for these five moraines derived from the range

of growth rates span from 860 cal yr BP based on the slow growth curve (Brooks Range)

to 20 cal yr BP using the fast growth curve (Kenai Peninsula, Figure 11). The moderate

growth curve from the Seward Peninsula (with a ~30-yr colonization time, see below)

yields an average age of 90 cal yr BP (~1860 AD) for these LIA maximum moraines.

Although the control points from the Seward Peninsula growth curve are not well-

constrained around this time period, the Alaska Range growth curve yields a similar age

for this lichen size. Because lichens date the stabilization rather than the initial

deposition of a moraine, this age is a minimum estimate of the LIA maximum in the

Cascade River valley.

Two of the lichen stations (12, 13) are up-valley of the LIA terminal moraine of

the Cascade Glacier and were used to calculate a retreat rate. Lichens were measured on

stable boulders overlying bedrock that were deposited upon ice retreat from the site.

Page 29: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

21

Station 12 is located 990 m from the 2004 ice margin and was dated to 1945 AD using

the Seward Peninsula growth curve (with the 17-yr colonization time of Solomina and

Calkin (2003); Table 4). Station 13 is 540 m from the terminal moraine with an age of

1965 AD, but this age is too young as the site lies ~30 m beyond the 1957 ice front. This

implies that the 17-yr colonization time built into the Seward Peninsula growth curve

may be too short for the Cascade Lake area. A longer colonization time of about 30-yr is

also indicated by the lack of lichens on surfaces exposed after 1984 and by only the

earliest stages (<1 mm) of lichen colonization on the surfaces exposed between 1957 and

1984. The lichen ages reported in this thesis are therefore adjusted to be ten years

younger than the ages derived from the Seward Peninsula growth curve.

The rate of retreat at Cascade Glacier was reconstructed by constraining the ice

margin of Cascade Glacier at five points, using lichen ages and aerial photographs,

namely: the 2004 observed ice front, 1984 and 1957 aerial photographs, lichen station 12

with an age of 1935, and the maximum LIA moraine dated to ~1860 AD based on

lichenometry. These points indicate that Cascade Glacier retreated at the following

average rates: 1860-1935 = 5 m/yr; 1935-1957 = 21 m/yr; 1957-1984 = 8 m/yr; 1984-

2004 = 16 m/yr (Figure 12). The overall retreat rate for Cascade Glacier since the LIA

maximum was 11 m/yr.

The retreat rate of Cascade Glacier from its maximum LIA extent was not steady

over the last 140 yr, reflecting climatic variations and topographic constraints. The

lowest retreat rate was documented between 1860 and 1935 (5 m/yr). This time period

includes the climate recovery following the termination of the LIA. The shift to a rapid

retreat between 1935 and 1957 (21 m/yr) is more difficult to explain. The period of

Page 30: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

22

1947-1976 coincides with a ‘cool’ PDO and weakened AL, which would have produced

cooler conditions in southwest Alaska (Papineau, 2001). These conditions should have

inhibited glacier retreat, whereas the warmer temperatures associated with the preceding

‘warm’ PDO (strengthened AL) period of 1925-1946 should have promoted glacier

retreat. Glaciers respond more rapidly to the affects of summer ablation on the glacier

toe than they do to winter accumulation during positive mass-balance years. Because the

response time of mountain glaciers averages 10-15 yr (Porter, 1986), the retreat rate of

Cascade Glacier between 1946 and 1957 may be out-of-phase with the PDO and AL and

may not have yet responded to the positive mass-balance years of 1947-1976. The 1957

and 1984 aerial photographs show the terminus of Cascade Glacier at nearly the same

location in a bedrock-constrained, narrow section of the valley, which explains the lower

retreat rate (8 m/yr) during this period. The most recent shift (post-1984) in retreat rates

to 16 m/yr is likely the glacier response to ‘warm’ PDO shift from 1977 to 1995 with

some lag in response time.

The multiple late Holocene advances recognized by Levy et al. (2004) at Waskey

Lake (~30 km south of Cascade Lake) contrast with the single LIA moraines recognized

in Cascade Lake study area. Both this study and Levy et al.’s (2004) used the same

lichenometric methods, yet Levy et al. (2004) identified three periods of ice advance at

900-1800, 350-1050, and 200-650 cal yr BP, none of which are represented in this study

area. Conversely, the most extensive advance in my study area took place at 90 cal yr

BP, but is not represented in Waskey Lake valley. This discrepancy in Neoglacial ice

advance is difficult to understand. The glaciers of both study areas are similar in size and

occupy mountains with similar physiography and climate regime. None of the glaciers

Page 31: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

23

overlie unconsolidated sediment that may lead to variations in sliding velocities; rather,

they directly overlie bedrock. The Waskey Glacier is underlain by granodiorite, whereas

the Cascade River valley glaciers overlie meta-sedimentary rocks with significant

regional faulting. It is a possibility that differences in sub-glacial hydrology (bedrock

permeability) may cause some of the variations in ice extent between the Cascade River

valley glaciers and Waskey Glacier.

The late 19th century age for the LIA moraines in Cascade River valley

corresponds to LIA moraine ages elsewhere in Alaska. In the northeastern Brooks

Range, for example, LIA retreat occurred around 1890 AD (Evinson et al., 1996). Late

19th century moraine stabilization has also been recognized in the southeastern coastal

glaciers, and the Kenai Mountains (Wiles and Calkin, 1994; Calkin et al., 2001). Taken

together, these ages suggest that the second half of the 19th century was marked by

widespread moraine formation across Alaska.

Equilibrium-Line Altitude. The modern glaciers (1957) in the area of Cascade Lake have

an average size of 1.6 km2 (n = 8) and an average ELA of 980 m ± 90 m (Table 5). The

glaciers in the study area extended 0.2 to 1.7 km farther down-valley during the LIA, an

average of 0.6 km (65%) larger than their 1957 counterparts. Chikuminuk Glacier

experienced the largest down-valley expansion of 1.7 km. Cascade Glacier’s ELA of 980

m coincides with the mean of the average ELA and its 1.4 km down-valley expansion

during the LIA is the second largest.

The average ∆ELA (lowering of ELA from LIA maximum to 1957) for the

Cascade Lake area was 23 ± 13 m (Table 5). ∆ELA ranges from 45 m (Cirque 2) to 5 m

Page 32: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

24

(Cirque 1). The difference in ∆ELA between these two cirque glaciers is unexpected

because they have similar southerly aspects, are within 1 km of one another, and have

similar modern ELAs (1010 and 1050 m respectively), but Cirque 1 contains the most

exposed glacier with an extensive, low-relief forefield, which may explain its limited

expansion during the LIA. ∆ELA does not appear to covary with glacier area, head

altitude, or aspect (Table 5).

The modern ELA of 980 ± 90 m determined in this study is similar to a GIS

assessment of 106 cirque glaciers in the Ahklun Mountains completed by Manley (1999).

Manley estimated a regional ELA of 929 ± 127 m, with 73% of the variance in ELA

accounted for by microclimate variables (winter precipitation, summer temperature,

aspect, etc.). Manley (1999) also demonstrated that the Ahklun Mountains cirque

glaciers are sensitive to climate changes that affect accumulation and ablation. For

example, a 50 m rise in the ELA would result in a 40% decrease in the total glacier area.

The only other ELA analysis from southwestern Alaska is that of Levy et al. (2004) for

Waskey Lake. Their ∆ELA estimate for the LIA (35 ± 22 m) overlaps with the Cascade

Lake study area. Additionally, Skiorski (2004) and Daigle and Kaufman (2006)

determined ∆ELA’s of 25 ± 17 m in the Brooks Range and 50 ± 20 m on the Kenai

Peninsula, respectively. In contrast, ∆ELA values reported for the LIA in Alaska by

Calkin et al. (1985) range from 100-200 m. Their calculation of ‘modern’ ELA was

based on mass-balance data from 1976-1982, whereas Levy et al. (2004) Skiorski (2004),

Daigle and Kaufman (2006), and my study used topographic maps and aerial photographs

to derive modern ELA values. The location of the ‘modern’ ice front is a cumulative

product of the previous decade’s (post-LIA) climate, whereas the mass-balance studies

Page 33: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

25

reflect the ELA of the limited study period. Therefore the higher ∆ELA values of Calkin

et al. (1985) may be attributed to the high ELA values of the 20th century.

The estimated ∆ELA in the Cascade Lake study area (23 ± 13 m) can be used to

estimate temperature lowering during the LIA. Using a typical environmental lapse rate

of 6°/km (Porter, 2001) and assuming no change in precipitation, results in a cooling of

0.13 ± 0.07 °C during the LIA. This lapse rate is only slightly lower than the regional

environmental lapse rate (7°/km) calculated from the summer temperature data in this

study. This amount of cooling is significantly less than the LIA cooling of 2-3°C

reported for the Brooks Range (Calkin et al., 1985), ≤ 1 °C temperature depression for the

Wrangell Saint Elias Mountains (Davi et al., 2003), ~1.7 °C cooling in the Alaska Range

(Hu et al., 2001), and the average 0.5-1.5 °C reported for the Northern Hemisphere

(Grove, 2003; Maasch et al., 2005). The significantly cooler temperatures of Calkin et al.

(1985) derive from their large ∆ELA values. The calculation of temperature cooling

from lapse rates and ∆ELA values is dependent on the assumption that no change in

moisture affected glacial expansion. However, if the northeastern Ahklun Mountains

were drier during the LIA, this would have limited glacier expansion despite cooler

temperatures. This scenario of a cold and dry LIA maximum in the study area may

explain the discrepancy in LIA temperature cooling calculated from lapse rate values.

Page 34: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

26

CHAPTER 4: CASCADE LAKE SEDIMENTS

Background

Cascade Lake was targeted for a lake sediment study because it has a high

sediment load and is hydraulically connected to up-valley glaciers that show evidence of

late Holocene fluctuations. Additionally, Waskey Lake (Levy et al., 2004) to the south,

and Arolik Lake (Kaufman et al., 2003) to the west have been used successfully to

reconstruct paleoclimates in the region. Cascade Lake sediments were analyzed for a

variety of proxy indicators including their geochronology, magnetic susceptibility (MS),

organic-matter content (OM), biogenic silica (BSi), and lamination stratigraphy.

In lake sediments, MS is affected by organic content, density, grain size and

mineralogy (Dearing, 1999), and it is not always possible to discern which property is

forcing the MS signal. For this reason, MS is used to identify shifts in sedimentation

patterns that can be correlated between cores and in some cases attributed to climate

variability when used in combination with other proxy records. Variations in OM reflect

a number of factors, including lake productivity, minerogenic flux, the input of terrestrial

vegetation to the lake and diagenetic preservation. Because climate fluctuations can

directly affect each of these factors, OM is a valuable indicator of climate (e.g., Willemse

and Tornqvist, 1999; Kaufman et al., 2003; Bakke et al., 2005). BSi reflects the

abundance of diatoms in lacustrine sediments. Diatoms are single-celled algae that

typically dominate internal productivity in a lake and are sensitive to environmental

changes, making them good climate proxies (e.g., Willemse and Tornqvist, 1999;

Anderson, 2000). The deposition of laminations in a lacustrine setting can be driven by a

Page 35: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

27

variety of hydrologic and meteorological controls (Leeman and Niessan, 1994;

Lamoureux, 1999). This relationship between climate and lamination thickness has been

used to produce quantitative and qualitative estimates of climate variability on annual to

seasonal scales (Hughen et al., 2000; Moore et al., 2001; Loso et al., 2006). Each of

these parameters were analyzed in Cascade Lake and interpreted for patterns that may be

indicative of climate variability.

Methods

Bathymetry, Acoustic Profiling, and Core-Site Selection. Core site selection was based on

bathymetric data and acoustic profiles. Bathymetric data were acquired with a Lowrance

sonar unit and subsequently plotted in GIS with the shoreline based on 1973 maps. The

sub-bottom stratigraphy was surveyed using an EdgeTech GeoStar acoustic profiler

operating between 4 and 24 kHz. This profiling identified regions of continuous and

condensed stratigraphy (avoiding slumps, methane, etc.). Three sites (distal-shallow,

proximal-shallow, and proximal-deep) were targeted for percussion coring (Figure 13).

Cores up to 6 m in length (7.5 cm diameter), were successfully recovered from these sites

(Table 6). Coring was completed from an anchored floating platform and was continued

until the core tubes were filled or an impenetrable layer was encountered. Because the

uppermost sediments in the long cores are often disturbed, surface cores were also

recovered from each site. These preserve the sediment-water interface and allow for an

evaluation of the stratigraphic completeness in the long cores. Surface cores were either

sub-sampled in the field into 0.5-cm increments or dewatered and packed. All cores were

shipped to Northern Arizona University and placed in cold storage.

Page 36: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

28

Lithostratigraphy. Cores were split and photographed at ~10-cm intervals to document

sediment color prior to oxidation. General sediment descriptions (color, layers, bedding

thickness, texture, etc.) were logged. Unique features, such as tephra beds (described

below), were sub-sampled and viewed under a binocular microscope.

Additionally, a test pit was dug on the northern shoreline of Cascade Lake

(Figures 13). The stratigraphy of the test pit was described and sub-samples were taken

for radiocarbon analysis and visual comparison to the lake core stratigraphy.

Geochronology. A variety of methods were employed to determine the ages of the

lacustrine sediments, each with well-established practices. These included: radiometric

210Pb and 137Cs dating, tephrochronology, and radiocarbon dating. 210Pb analysis was

completed on surface core 02A at the University of Illinois by alpha counting and on

Core 02 at University of Southern California by gamma spectrometry in conjunction with

137Cs analysis. Contiguous 0.5-cm-thick (~1 cm3) samples were submitted for analysis.

Both 210Pb profiles were modeled using a constant rate of supply (CRS) model because

this model copes with variable sedimentation rates, which probably fluctuated

significantly (Appleby, 2004).

Tephras were located by their coarser grain size and their volcanic glass shards

and pumice. MS peaks were also targeted due to the association of volcanic glass and

magnetite. Each tephra was described under a binocular microscope to compare with

other tephras known from the region. I analyzed the major-element geochemistry of the

most prominent (visible and thickest) tephra in the Cascade Lake cores to correlate it with

Page 37: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

29

tephra of a known age. The tephra was sampled from the center of the unit in Core 02 at

a depth of 2.85 m. It was cleaned with 2% NaOCl and dried. Glass grains were hand

picked under a binocular scope, epoxied onto a glass slide, polished with 1000 µm grit,

then carbon coated for analysis on a Cameca MBX electron microprobe at Northern

Arizona University. A single point on each of 31 glass grains was analyzed along with a

standard of basaltic glass (USM 11349811) used to calibrate the microprobe prior to use.

Total oxide percents of 10 elements were normalized to 100%. Similarity coefficients

(Borchardt et al., 1972) were determined by comparison with published values for the

3500 cal yr BP Aniakchak tephra (Riehle et al., 1987; Béget et al., 1992; Levy, 2002).

Lake sediment samples were sieved every ~5 cm down the core profile in search

of macrofossils, but no sample contained sufficient material for radiocarbon analysis. A

single radiocarbon age was obtained from a buried soil in the test pit on the north

shoreline of Cascade Lake. A sub-sample of the buried soil was wet-sieved and

handpicked for vegetation macrofossils. Clean and dry macrofossils were analyzed by

AMS at the University of California, Irvine. The radiocarbon age was calibrated using

CALIB v. 5.0 (Stuvier et al., 1998) and the reported age is the median probability age of

the 1σ calibrated age range.

Magnetic Susceptibility. MS was measured on the split core face using a Bartington MS2

meter, a hand-held MS2E1 detector, and Multisus computer software to correct for

measurement drift. Refrigerated cores were brought to room temperature before analysis.

Measurements were taken every 0.5 cm with drift correction every five readings. MS

values are reported in Systéme International (SI) units and were taken at a 0.1 sensitivity

Page 38: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

30

range setting. Sediments with anomalously high MS values were examined under a

binocular microscope to confirm the presence of tephra shards, or unique mineralogy.

Organic-Matter Content. OM and bulk density were measured on two of the three cores.

Contiguous 1 cm3 samples were taken every 1 cm down to a depth of 198 cm in Core 01,

and over the entire length of Core 02. In a few places sample resolution was reduced

over coarse-grained, dense tephra units. Samples were processed using the procedure

described by Dean (1974). They were dried at 90°C for 12 hr, and then weighed to

measure dry bulk density. OM was approximated by LOI after combusting samples at

500°C for 5 hr. Because inorganic carbon content is negligible in the sediments, it was

not measured.

Lamination Stratigraphy. Although laminations were faintly visible on much of the

surface of the cores they were too diffuse for accurate measurement of thickness based on

visual inspection. Therefore, the core face was sub-sampled to a depth of 140 cm and

imaged by X-radiography. Slabs were 15 x 0.4 cm and were taken from the core axis

with ~1 cm of overlap between the slabs. These uniform-thickness slabs were X-rayed at

the NAU health center on detail film (3/20 sec, 50 mA, 50 kVp). Laminations were

visible on the X-rays as discrete zones of millimeter-scale, alternating high and low

density. Images were digitally spliced and lamination couplets were counted (four times)

and the thickness of both the high- and low-density layers and the total couplet was

measured (Figure 14).

Page 39: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

31

Biogenic Silica. Biogenic silica (BSi) analysis was performed using the methods

described by Mortlock and Froelich (1989). Contiguous 2-mm-thick samples were taken

from Core 02 between 0 and 0.15 m, and from 0.15 to 1.50 m at a 5 mm sample interval.

The rest of the core was sampled in 5-mm-thick sections at 5-cm intervals. Hydrogen

peroxide (H2O2) and hydrochloric acid (HCl) were added to ~100 mg of dried and

powdered sample to remove organic material and carbonate. BSi was then extracted with

10% sodium carbonate solution (Na2CO3). The supernate containing dissolved silica was

then treated for molybdate-blue spectrophotometry through the reduction of the

silicomolybdate complex. BSi concentration was determined with a HACH DR/2000

spectrophotometer set to 812 µm wavelength and zeroed against a blank standard of

Na2CO3. Absorbance values were converted to BSi content using a standard curve

created by NAU lab technician Caleb Schiff (Fall 2005).

Samples were run in groups of 40 and experienced some batch effects

predominantly caused by slight variations in the mixing of the siliciomolybdate reduction

solutions. In order to adjust for these batch effects, three samples from the initial batch

were treated as standards and processed with every subsequent batch. Batches were then

calibrated to the original batch by adjusting BSi values by the average difference of the

three sample standards. Although the reported BSi content depends on these adjustments,

the relative difference among sampled is reproducible to ± 0.3% (1σ) based on multiple

runs of the standard samples.

Page 40: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

32

Results and Discussion Bathymetry, Acoustic Profiling, and Core Sites. Cascade Lake contains a deep trough

(>50 m) running from the main inflow on the south end of the lake to the outlet on the

north. Broad, sloping deltas extend from the main and eastern inflows (Figure 13).

Acoustic profiling revealed generally continuous stratigraphy with localized zones of gas

disruption (Figure 15). The recovered cores did not penetrate the full stratigraphic

sequence visible in the acoustic profiles, with about 3 m of sediment remaining below the

base Core 01; however, there was little acoustic stratigraphy visible beneath the base of

Core 02. This could be because the stratigraphic sequence of this core is more complete

or because the acoustic signal was degraded.

Lithostratigraphy. Core 01, from the north distal core site, consists of five main

stratigraphic units with gradational contacts (Figure 16). (1) The basal unit (4.7 to 4.2 m)

is composed of sand interbedded with gray (5B 5/1) mud. The fine-grained layers are

separated by sharp contacts from the sandy mud. Unit 2 (4.2 to 3.6 m) is well-laminated

with laminations apparently defined by grain size (silt rich vs. clay rich). The

laminations range between ~0.25 and 1.0 cm thick. Unit 3 (3.6 to 2.8 m) consists of

brown (2.5Y 5/4, 106 5/6) clay-rich mud. Unit 4 (2.8 to 1.1 m) is blue-gray clay-rich mud

with three coarse-grained tephras layers. Unit 5 (1.1 to 0 m) is light-brown (2.5Y 5/4, 5Y

5/1) mud with very faint laminations in the uppermost ~30 cm.

The lithostratigraphy of Core 02 is very similar to Core 01 (Figure 16). It

comprises five fine-grained units defined by color differences and separated by

gradational contacts. (1) The basal unit (4.4 to 3.8 m) consists of sand interbedded with

Page 41: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

33

gray, clay-rich mud. The sand layers are better defined than those of Core 01. Unit 2

(3.8 to 3.7 m) contains well-defined laminations > 1 cm thick, separated by sharp

contacts. Unit 3 (3.7 to 3.1 m) is composed of clay-rich, light-yellow-brown (10YR 5/6)

sediment with coarser-grained (not tephra) beds averaging 2 cm thick. Unit 4 (3.1 to 1.6

m) consists of blue and olive gray, fine-grained sediment (5B 5/1, 5Y 5/2) with five

coarse-grained tephras. Unit 5 (1.6 to 0 m) is olive to yellow-brown (2.5Y 4/4, 10YR

5/6, 2.5Y 6/2) mud with faint laminations throughout. These laminations are a centimeter

to millimeter thick.

Core 03 (deep, main basin core) is markedly different than Cores 01 and 02

(Figure 16). It contains well-defined, thick (0.5-3.0 cm) beds of alternating silt- to clay-

rich mud and disturbed and deformed sediments that interrupt the bedded stratigraphy.

Unit 1 (6.0 to 4.4 m) is composed of deformed sediments, coarse-grained beds and > 1-

cm-thick laminations. Unit 2 (4.4 to 3.2 m) consists of coarse-grained layers interbedded

with massive mud. Unit 3 (3.2 to 0.9 m) is predominantly deformed sediments with some

discrete units of mud. Unit 4 (0.9 to 0 m) is composed of well-defined laminations that

are discontinuous and separated by deformed sediments. Overall, Core 03 contains four

zones of well-formed, grain-size-defined laminations. No tephras were observed in the

Core 03.

The tops of Cores 01 and 02 were correlated with their sister surface cores using a

distinct fine-grained marker bed that grades upward from dark brown to red-brown. In

Core 01 this marker bed is at 8.0-8.5 cm depth, and at 8.5-9.0 cm in the surface core. In

Core 02 the bed is thicker and at a depth of 13.5-15.0 cm compared to 14.0-15.0 cm in

the surface core (Figure 17). Using these correlations, ~0.5 cm is missing from the

Page 42: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

34

surface of each of the long cores and core-tube depths were adjusted accordingly so that

all reported depths are relative to the lake floor.

The lithostratigraphy of Cores 01 and 02 can be roughly correlated (Figure 16).

Both contain a thick (~15 cm) tephra (at 2.12 m in Core 01 and 2.85 m in Core 02), and

the basal sediments are similar in both cores. Laminations are better preserved in Core

02, though they are subtly visible on the surface of Core 01.

Because discrete horizontal beds are preserved in sections of Core 03, the

disturbance and deformation probably was not caused by the coring process. Instead, the

sediment deformation is likely caused by underflow currents or slumps within the deep

trough where the core was recovered. The thick tephra unit visible in Cores 01 and 02 at

2.12 and 2.85 m, respectively, was not found in Core 03. This, in combination with the

thick laminations of Core 03 (> 1 cm), indicates that sedimentation rate at the Core 03

site is significantly higher than that of Cores 01 and 02. For these reasons little further

work was completed on Core 03.

I focused most of my work on Core 02 because the laminations were better

preserved than in Core 01, likely due to its closer proximity to the main inflow, and it has

a longer sedimentary sequence and is less disturbed than that of Core 03. The proximal-

shallow water location of Core 02 also might be more sensitive to climate variability due

to the greater influence of the main inflow compared to the distal site of Core 01.

The test pit dug on the northern shoreline of Cascade Lake was 25 cm deep and

was subdivided into five units (Figure 18). (1) The base of the soil pit ended in a poorly

sorted, sub-angular to sub-rounded cobble gravel. This is overlain by Unit 2 (24.5-10.0

cm) a beige-orange to brown-orange tephra that grades from coarse to fine sand and

Page 43: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

35

contains pumice shards visible with a hand lens. Unit (3) (10.0 to 4.3 cm) is a buried A-

horizon that is dark-brown to black. Unit 4 (4.3 to 2.6 cm) contains a discontinuous gray

tephra, with pumice lapilli, forming discrete lenses and mixed with organic material from

the buried soil below. Unit 5 (0-2.6 cm) is organic matter of living tundra.

210Pb and 137Cs. The two 210Pb profiles analyzed from cores 02 and 02A are quite similar

(Figure 19, Table A-1). In the core 02A, excess 210Pb activity levels off at 1.0 dpm/g at

~13 cm depth, resulting in an age of 1856 AD at 12.25 cm (CRS model) and an average

sediment flux of 0.08 g/cm2/yr. The profile indicates a shift in sediment accumulation

rate at ~8.25 cm (1948 AD). Below this point the sedimentation rate averages 0.48

mm/yr, compared with 1.76 mm/yr above it. This sedimentation-rate shift is only

constrained by four points at the base of the profile. The 210Pb profile from Core 02

results in an age of 1851 AD at a depth of 12.5 cm (Figure 19) similar to Core 02A. The

210Pb activity of Core 02 exhibits a shift in sedimentation rate from 0.40 mm/yr below

8.25 cm to 1.48 mm/yr above this point. This shift occurred a 1952 AD according to the

210Pb age model from Core 02, but is constrained by only two points. The 210Pb profile of

Core 02 was supplemented with 137Cs analyses from the same sediments. 137Cs peaked

with atmospheric nuclear testing in 1963, and therefore can be used as an independent

measure of sediment age. The activity of 137Cs shows a prominent peak from 8-6 cm

with a maximum at 6.25 cm (Figure 19). According to the 210Pb profile, a depth of 6.25

corresponds to an age of 1971, not 1963.

The two 210Pb models for core 02A and Core 02 differ slightly, especially in the

youngest sediments, but are similar in sediments older than the 1800’s. Both cores

Page 44: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

36

exhibit a major shift in sedimentation rate at 1950 AD. Because the 210Pb profile produce

by gamma spectrometry was analyzed on sediments directly from Core 02, which was

also analyzed for other parameters, I focused on this profile and incorporated it into the

age model. The offset between the 210Pb and 137Cs ages determined from the same

sediments might suggest an age error in the 210Pb ages. However, 137Cs was only

analyzed every 0.5 cm, and the age of 1963 lies between the two points that define the

137Cs peak in this profile. Assuming that the 137Cs peak occurs at ~6 cm, then the 210Pb

ages are within ~4 yr.

Radiocarbon. A single 14C age was obtained from the top 1 cm of the buried soil in the

test pit on the lakeshore. The sample consisted of a mixture of detrital wood and leaf

fragments (UCI AMS# 23856). The sample yielded a radiocarbon age of 435 ± 15 cal yr

BP, which calibrates to 505 ± 20 cal yr BP. This is a maximum age for the tephra that

overlies this buried soil.

Tephrochronology. Five prominent tephras were recognized in Core 02 at 3.41, 3.33,

3.24, 2.85, and 2.59 m (Table 7, Figure 16). All of these tephras were also recognized in

Core 01 (except the uppermost tephra). In addition, two tephras were described from the

test pit dug on the northern lakeshore (Table 7, Figure 13).

Two distinct populations were recognized in the geochemical data from the 2.85

m tephra in Core 02 (Figure 20, Table A-2), one with 74 ± 3% SiO2 (n = 16) and one with

59 ± 4% SiO2 (n = 15). This bimodal composition could derive from either zonation

within the magma chamber or from the mixing of multiple eruptions into this single

Page 45: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

37

layer. Similarity coefficients between these average data and published geochemical data

for the Aniakchak tephra are as follows: 0.75 with Riehle et al. (1987, for the high SiO2

group); 0.76 with Béget et al. (1992, for the high SiO2 group); and 0.75 with Levy (2002,

both SiO2 groups). However, four of the individual point analyses showed a significant

(>0.95 similarity coefficient) correlation with the Riehle et al. (1987) samples.

The geochemistry of the 2.85 m tephra correlates poorly with the reference

samples of the 3500 cal yr BP Aniakchak tephra. On the other hand, the bimodal

composition of this tephra is indicative of the Aniakchak tephra. Of the four tephras

located in Ahklun Mountain lakes and analyzed for geochemistry by Levy (2002), only

the Aniakchak tephra exhibits this bimodal composition. The Aniakchak tephra is also

the thickest and coarsest-grained tephra in the region (Kaufman et al., 2003; Levy et al.,

2004), as is the 2.85 m tephra in Cascade Lake. Scanning electron microscope images of

the 2.85 m ash from Cascade Lake and a sample of known Aniakchak tephra also show

similar shard morphologies (Figure 21). Both samples have fibrous pumice, blocky

clasts, and bubble-walled platelets, all of which were also identified by Riehle et al.

(1987) in their distal Aniakchak samples. For these reasons, I believe that the 2.85 m

tephra in Core 02 is the Aniakchak tephra (= 3500 cal yr BP).

The stratigraphic position of the Aniakchak tephra helps to correlate other

tephras. Assuming the Aniakchak tephra is correctly identified in the Core 02, then the

base of the core is ~11,000 cal yr BP based on the extrapolation of sedimentation rates.

Waskey Lake cores contain four other tephras, only one of which is younger than 11,000

cal yr BP (Levy et al., 2004) and therefore young enough to be in the Cascade Lake

cores. This tephra, Tephra B, is a “pure, 1-cm-thick, dark tephra…dominantly fine sand

Page 46: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

38

to coarse silt…the glass shards are platy, brown, phenocrystic and microlitic…bubble-

wall junctions are present”. In both Waskey Lake and Arolik Lake (Kaufman et al.,

2003), Tephra B dates to ~6100 cal yr BP, and is the next recognized tephra below the

Aniakchak tephra. In the Cascade Lake cores, the 3.24 m tephra is the next lower and a

likely candidate for Tephra B. However, the tephras at 3.33 and 3.41 m also fit the

description of Tephra B and are possible correlatives, although none were analyzed for

major-element geochemistry.

Overall, more Holocene tephras are observed in the Cascade Lake cores than in

any other lake core from the Ahklun Mountains; however, the record is complicated. The

505 cal yr BP gray tephra in the test pit could not be located in the cores despite an

intensive search. Additionally, the tephra at 2.59 m is located only in Core 02, and was

not found in the test pit or in Core 01. Of the other lakes cored in the Ahklun Mountains,

only Little Swift Lake (Axford and Kaufman, 2004) contains two tephras older than the

Aniakchak tephra and younger than 12,000 cal yr BP. Both Arolik Lake (Kaufman et al.,

2003) and Waskey Lake (Levy et al., 2004) contain a single tephra over this interval.

This is inconsistent with the three tephras in Cascade Lake below the Aniakchak tephra.

The inconsistencies among these tephra records are difficult to explain. The deposition

of tephra onto a snow-covered landscape and ice-covered lake may retard deposition of

the tephra in an individual lake. Additionally, spatial patchiness has been documented in

the physical parameters of lake sediments (e.g., water content, organic-matter, pigment,

and phosphorus; Downing and Rath, 1988). Similar spatial heterogeneity might apply to

the deposition of tephra in lake sediments. The presence of the 2.59 m tephra in Core 02

but not in Core 01 indicates that tephras are not deposited uniformly within large lakes.

Page 47: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

39

Age Model. An age model was constructed for Core 02 based on the 210Pb CRS model

and the correlated ages of the Aniakchak tephra and Tephra B. Because Tephra B was

not definitively located in Core 02 and could be any of the three tephras between the 3.24

or 3.41 m, two age models were created to bracket the range of these possible

correlations (Figure 22). The age model was subdivided into three segments. The basal

section is an assumed linear sedimentation between Tephra B and the base of the

Aniakchak tephra. The average of the Tephra B alternatives results in a sedimentation

rate of 0.2 mm/yr. The deposition of the Aniakchak tephra between 2.85 and 2.70 m was

interpreted as being instantaneous. The next segment between the Aniakchak tephra and

the 210Pb age at 12 cm was assumed to be linear, with a sedimentation rate of 0.8 mm/yr.

For the upper segment, the CRS 210Pb model was approximated by a 3rd-order polynomial

fit with increasing sedimentation rates upwards (from 0.4 mm/yr at the bottom to 1.5

mm/yr at the top).

The pronounced shifts in sedimentation rates can be reconciled with other

evidence. The factor-of-four lower sedimentation rate below the Aniakchak tephra

agrees with the sedimentation-rate shift that Levy (2002) documented in Waskey Lake.

The striking increase in sedimentation rate at the top of the core (post-1950) corresponds

to the high (21 m/yr) retreat rate of Cascade Glacier from 1935-1957 and a period of

accelerated warming across the arctic (Overpeck et al., 1997).

If the low sedimentation rate between the Aniakchak tephra and Tephra B is

assumed to continue to the base of the core, then the basal sediments are approximately

11,000 cal yr BP using an average linear sedimentation rate of 0.2 mm/yr. This

continuation of a linear, low sedimentation rate below Tephra B is similar to that

Page 48: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

40

determined by Levy et al. (2004) for Waskey Lake and the basal age corresponds with the

a northwestern Ahklun Mountain deglaciation age of ~12,000 cal yr BP (Manley et al.,

2001).

Magnetic Susceptibility. The average background MS value in all three cores is ~30 x 10-

6 SI (Figure 23, Table A-3). Both Core 01 and 02 show elevated MS levels in the basal 1

m and discrete zones of high MS (> 50 x 10-6 SI). In Core 01, the peaks in MS occur at:

3.29 m (75.5 x 10-6 SI), 2.73 m (175 x 10-6 SI), 2.27 m (58 x 10-6 SI), and 2.1 m (412 x

10-6 SI). The sediment at 2.1 and 2.73 m contains tephra. There was no visible change in

sediment composition at the 2.27 m MS peak, but the MS peak at 3.29 m contains >80%

diatoms (primarily Cyclotella). MS peaks in Core 02 occur at: 3.35 m (105 x 10-6 SI),

2.84 m (343 x 10-6 SI), and 2.59 m (114 x 10-6 SI). The peaks at 2.59 m and 2.84 m

correspond with tephras (see Tephrochronology). The sediments at 3.35 m (6250 cal yr

BP) are similar to those at 3.29 m in Core 01; both contain a high percentage of

Cyclotella diatoms. MS peaks in Core 03 are all less than 50 x 10-6 SI, and values vary

between 7 and 47 x 10-6 SI.

Most of the MS peaks in the Cascade Lake cores coincide with tephras. The

presence of multi-domain magnetite in coarse-grained tephras yields an elevated MS

value when compared to the ambient lake sediments. The only MS peaks that do not

coincide with tephra are the ~1-cm-thick layers containing >80% diatoms at 3.29 m in

Core 01 and 3.35 m in Core 02. The association of elevated MS values with high diatom

concentrations is unexpected because diatoms are pure opaline silica and should not have

high MS. Possibly, the algal bloom responsible for the high concentration of diatoms

Page 49: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

41

was also associated with increased growth of magnetobacteria. Magnetobacteria have

been shown to respond directly to the same environmental factors that affect general lake

productivity (Kim et al., 2005). Additionally, the draw-down of oxygen by a large algae

bloom could produce anoxic conditions favorable for magnetobacteria growth (Kim et

al., 2005). The general rise in background MS at the base of Cores 01 and 02 is related to

the lithologic variability of sandy mud interbedded with fine-grained beds. This rise

could reflect the mafic-rich sandy beds or the increased density of the fine-grained beds.

Organic-Matter Content. OM in Core 01 ranges from 3.1 to 5.0%, whereas in Core 02 it

varies between 1.4 and 4.7% (Figure 24, Table A-4). Bulk density ranges between 0.7 to

2.2, and 0.5 to 2.1 g/cm3, respectively, in the two cores. The OM profile for Core 02 has

relatively low OM values at the base of the core that rise rapidly to a broad peak at ~3.55

m (7250 cal yr BP); then OM values gradually decrease to the top of the core. Bulk

density shows little overall change in this core. OM and bulk density generally decrease

towards the core top in Core 01. The highest OM values are at 1.54 m (4.9%) and 0.9 m

(4.8%). Each of these peaks is constrained by multiple samples.

Bulk density and OM in Core 02 are weakly correlated (r = -0.42; p = 1.3 x 10-18).

The overall low OM values reflect the pervasive minerogenic input from the glacial

inflows and the relatively low lake productivity and vegetation growth on the surrounding

landscape. The inverse relationship between OM and bulk density probably is due to

increased sediment accumulation (high bulk density) diluting the organic matter. The

decrease in OM in Core 02 at 3.55 m occurs over the prominent lithologic change from

gray mud to yellow-brown mud with well-developed, thick laminations. The lowest OM

Page 50: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

42

values at the base of Core 02 and at 2.85 m correlate with the coarsest sediments related

to a tephra and sandy mud in these intervals.

Laminations. A total of 316 lamination couplets were counted between 0.05 (core top)

and 1.45 m. These couplets averaged 4.2 mm in thickness, with the dense layers

averaging 1.8 mm and the less-dense layers 2.4 mm thick (Figure 25). Based on

observations from the core face, the dense (light) layers are associated with an increased

proportion of fine-grained (clay) sediment relative to the less-dense (dark) layers. Total

couplet thickness is more strongly related to the thickness of the less-dense layers (r =

0.88) than to the dense layers (r = 0.73). The thickness of the dense and less-dense layers

show no relationship to one another (r = 0.32). Lamination thickness is highly variable

ranging from 1 to 15 mm.

The discernable seasonal signal in the sediment-trap material suggests that

Cascade Lake laminations are annual (varves). However, their relatively large average

thickness and the geochronology suggest that this is not the case. The high variability in

lamination thickness, with the lack of a systematic down-core pattern of fluctuation (i.e.,

no section contains laminations persistently above or below average thickness), indicate

that lamination thickness is not a reliable proxy for climate fluctuation in this lake.

Applying the age model to the lamination counts indicates that the average

lamination encompasses ~6 yr. This is on the scale of the sub-decadal climate-regime

shifts that affect southwestern Alaska (AL, ENSO; Papineau, 2001). Changes in North

Pacific climate regimes have been shown to influence fluvial dynamics and likely

lacustrine sedimentation. For example, in southeastern Alaska, Neal et al. (2002)

Page 51: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

43

demonstrated that ‘warm’ PDO regimes correlated with higher-discharge winters (more

precipitation falling as rain as opposed to snow) and ‘cool’ PDO’s were related to higher

summer discharge (precipitation stored as snow and released during the peak of the melt

season). This type of multi-annual shift in runoff may drive shifts in sediment storage

and therefore the formation of laminations in Cascade Lake. Hodgkins et al. (2003)

showed that proglacial outwash plains can serve as either net sinks or net sources of

sediment with the shifts in sediment storage driven by the runoff regime. This variation

in sediment storage (potentially caused by North Pacific climate regime shifts) may

explain the variations in sediment flux responsible for lamination thicknesses.

The deposition and preservation of laminations can also be caused by processes

within the lake. The occurrence of large magnitude underflows or short-lived

environmental events (e.g., storms) can dominate the sediment record of some lakes and

result in the deposition of non-annual laminations (Lamoureux and Gilbert, 2004). These

processes are often associated with sedimentary structures that can be related to turbidity

flows or slumps, none of which were observed in Cascade Lake. Rhythmic laminations

were also recognized by Lemmen et al. (1988) in a glacial-fed lake on Baffin Island.

They interpreted the laminations as being caused by changes in the rate of sediment input

to the lake and/or changes within the circulation of the lake. They also attributed the lack

of varves to low total sediment input to the lake and the extremely fine nature of the

sediment entering the lake. Cascade Lake has relatively high levels of summer

suspended load, though concentrations were never measured, but the particle size is small

(avg. 4 µm) and may result in a long sediment residence time in the water column and the

non-deposition of varves.

Page 52: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

44

Biogenic Silica. BSi in Core 02 averages 2% (i.e. 20 mg/g) and ranges from 0.6 to 6%,

and shows significant variability including three prominent peaks (Figure 26, Table A-5).

BSi values are the lowest at the base of the core and rise sharply at 3.35 m (6.0%; 6275

cal yr BP). Between 3.35 and 2.65 m, BSi values decrease again (3.0%) before rising to

form a broad peak at 2.65 m (5.4%; 3500 cal yr BP). BSi values return to an average

1.6% and then peak between 1.18 and 1.03 m at 2.5%. The average BSi value towards

the top of the core (1.03 to 0.0 m) is 1.9%.

The sharp peak in BSi at 3.35 m encompasses the ~1-cm-thick layer of diatoms.

The deposition of these diatoms was probably short-lived (perhaps within a single

season) and not caused by a long-term climate fluctuation. The BSi peak at 2.65 m

occurs above the thick Aniakchak tephra deposited at 3500 cal yr BP. The deposition of

tephra within a lake can encourage algal blooms, especially in environments where Si is a

limiting factor (Telford et al., 2004). Although BSi was increasing just prior to the

deposition of the 2.85 m tephra, the peak is significantly higher than the extrapolation of

this rise, and the BSi recovers to a lower level within 0.4 m of the tephra, suggesting that

this spike in BSi was caused by the tephra deposition.

Page 53: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

45

CHAPTER 5: BIOGENIC SILICA AND CLIMATE INTERPRETATIONS

Diatoms are excellent climate indicators because they have a short life span and

respond directly and sensitively to physical, chemical and biological changes to their

environment. Because the translation of climate fluctuations to diatom productivity is

indirect, several assumptions must be made in order to use BSi as a proxy for climate,

among these is the assumption that BSi concentrations are directly related to the

productivity of diatoms.

The diversity and abundance of a diatom community in a lake is dependent on a

variety of factors, including: temperature, nutrient load, length of ice cover, turbidity,

thermal stratification, and duration of spring and fall turnover (Lotter et al., 1999;

Anderson, 2000; Smol and Cumming, 2000; Douglas et al., 2004). Each of these

environmental factors is directly linked to meteorological forcing (temperature, wind,

precipitation, etc.), and, in turn, to climate processes. The relationship between diatom

productivity and the physical/chemical characteristics of lakes is not always straight

forward. For example, Bradbury et al. (2002) described the relationship between spring

temperature and diatom concentration for Elk Lake, Minnesota. They discovered that

cold springs could result in either a small diatom population through late ice-out and

rapid over-turning or a large diatom population as a result of a slow transition to summer

and a long over-turning period.

Nutrient load is perhaps the most complex factor affecting diatom growth.

Silicon, nitrogen, and phosphorous are the dominant nutrients affecting the growth of

diatom communities and can be derived from a variety of sources. In arctic and alpine

Page 54: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

46

environments, silicon is primarily derived from groundwater and the dissolution of

diatoms at the sediment/water interface, nitrogen (N) enters a lake from soils and

vegetation, and phosphorous is regenerated from the lake sediment during the winter and

transported to the photic zone during spring circulation, but is ultimately supplied from

bedrock and eolian dust (Bradbury et al., 2002). The establishment and concentration of

alder (Alnus) within the catchment following deglaciation, strongly controls N

concentrations in southern Alaska lakes (Hu et al., 2001; Engstrom and Fritz, 2006). The

development of soil and vegetation therefore is a primary influence on diatom growth.

The connection between lake productivity and climate is further complicated by

the methods used to determine lake productivity. The use of BSi as a measure of lake

productivity is dependent on the assumptions that: (1) The BSi preserved in the lake

sediment is from algal growth rather than phytoliths, (2) there is little recycling of silica

(dissolution of diatoms at the sediment/water interface), (3) there is minimal grazing of

diatoms by higher trophic levels, (4) all algal species contributing to primary productivity

are preserved in the lake sediment, and (5) there is no significant change in the size and

silica concentration of the diatom species (Bradbury et al., 2002; Colman and Bratton,

2003). The validity of these assumptions is difficult to assess from the sediment record.

In studies of lake productivity from Lake Baikal some have proven to be inadequate. For

example, Qui et al. (1993) and Swann and Mackay (2006) both measured low BSi

content in sediment with high diatom concentrations based on point counts and other

techniques. This discrepancy was attributed to dissolution of the diatoms at the

sediment/water interface during periods of low sedimentation.

Page 55: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

47

Variations in BSi content in Cascade Lake are believed to reflect fluctuations in

lake productivity. The lake is ice-free for approximately five months of the year, and

likely has a significant nutrient load entering the lake, enabling a seasonal diatom bloom.

The rather high sediment load of the lake may limit the area of the photic zone available

for diatom growth, but surface area of the lake is large enough to support a significant

diatom population. Diatom dissolution is dependent on the pH, temperature, salinity and

ionic composition of the water and sediments (Ryves et al., 2001). Temperature was the

only one of these parameters measured in the Cascade Lake; however, it is unlikely that

extreme values of pH, salinity or ionic composition would be maintained in the lake for a

significant period of time, and diatom dissolution is considered to be in the lake. The

pronounced difference in BSi content in the trapped Cascade Lake sediments between

2004-2005 and 2005-2006 indicates that BSi is sensitive to interannual differences in

weather conditions and therefore to longer term climate relationships. Because the

laminations in Cascade Lake were determined to not be annual, I focused on fluctuations

in BSi as the primary indicator of paleoclimate variability.

Methods

In order to determine which climate parameter is most strongly correlated with

BSi in Cascade Lake sediments, I compared the BSi flux with meteorological records.

The 3rd-order polynomial fit to the CRS age model was used to infer the ages of the high-

resolution BSi samples analyzed from the top (0-14 cm) of Core 02 and to interpolate the

sedimentation rates of each 0.2-cm-thick BSi sample. The average temporal resolution of

the samples was determined to be 3 yr. BSi values were converted to BSi flux (BSi % x

bulk density x sedimentation rate) using bulk density values that were linearly

Page 56: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

48

interpolated between the 1-cm-thick samples analyzed for OM. This conversion to flux

accounts for the effects of variable sediment dilution. These flux values were then

correlated with meteorological parameters (e.g., mean annual temperature, June-July-

August average temperature) from the instrumental records at Dillingham (NCDC, 2005;

WRCC, 2006), discharge data from Nuyakuk River (70 km to the southeast; NWIS,

2006), and climate indexes from the North Pacific Ocean (e.g., North Pacific Index,

Aleutian Low Pacific Index; Ebbesmeyer et al., 1995; Mantua et al., 1997; Beamish et al.,

1997; NCAR, 2006). Correlations were made using both unsmoothed data and five-year

running averages of the climate data. The five-year running average was to account for

possible error in the 210Pb age model, smoothing in the sedimentary record (bioturbation),

and the prolonged influence of sub-decadal climate variability leading to ‘memory

effects’ (e.g., a rise in the groundwater table that sensitizes a landscape to runoff in the

years following). Temperature and precipitation data are discontinuous from the

Dillingham station with significant gaps in the 1970’s and 1930-1940’s. Temperature

averages were not used if more than three days were missing during a month and annual

precipitation values were acceptable only for years with no missing values (n-values in

Table 8).

Results

Correlations with BSi flux were made with 33 meteorological and North Pacific

climate parameters. Of these correlations, 11 were deemed of interpretive value (Table

8). BSi flux showed the highest correlation with Nuyakuk River discharge (r = 0.73, p =

2.2 x 10-6); the five-yr running average of mean May temperature (r = 0.70, p =9.16 x 10-

Page 57: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

49

5); and the five-yr running average of the Aleutian Low Pacific Index (ALPI, r = 0.68, p =

1.37 x 10-7). BSi flux also correlated with the Pacific Decadal Oscillation Index (PDO, r

= 0.42, p =0.003), and the Pacific Northwest Index (PNI, r = 0.41, p = 0.003), but not

with the North Pacific Index (NPI, r = 0.17, p = 0.229).

Discussion

Controls on BSi in Cascade Lake. I attribute the relatively strong correlations between

BSi, Nuyakuk River discharge, and May temperatures to the hydroclimate and landscape

processes that control nutrient input to the lake. Because Cascade Lake does not support

an anadromous fish population, all nutrients must enter the lake from terrestrial sources

(fluvial, eolian, through-flow, groundwater, etc.). The majority of fluvial input to

Cascade Lake traverses the glacial outwash plain of Cascade Glacier with the exception

of the small western inflow (Figure 13). Glacial runoff is typically low in nutrient

content (Qui et al., 1993), especially during periods of glacier retreat when the rock-water

(ice) contact time is decreased (Lanfreniere and Sharp, 2005). Because of this, nutrient

input to Cascade Lake (and algal growth) is probably more strongly related to

precipitation than to glacier melt. Summer precipitation and spring snow-melt (winter

precipitation) infiltrate the soil and flow over the landscape carrying nutrients to the lake.

The gage on the Nuyakuk River measures the outflow of Nuyakuk and Tikchik Lakes (70

km southeast of Cascade Lake). Because less than 0.01% of the drainage basin is

occupied by modern glaciers, summer ablation has little influence on Nuyakuk River

discharge. There is no precipitation data from Cascade Lake, and Dillingham is more

coastal than Cascade Lake, therefore Nuyakuk River discharge is the best approximation

Page 58: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

50

of precipitation at Cascade Lake. The influence of runoff on algal growth in a nutrient-

limited system has also been inferred for Grandfather Lake (Hu and Shemesh, 2003) and

Arolik Lake (Hu et al., 2003) in the Ahklun Mountains. While this terrestrial input

(related to precipitation) is likely the dominant source of Cascade Lake nutrients, some

nutrients are probably re-circulated within the lake and along the lake shoreline, driven

by wind mixing of sediments and erosion of the shoreline (e.g. Verschuren, 1999).

The relationship between May temperature and terrestrial runoff is more complex.

In 2004-2006, Cascade Lake experienced melt-out in the last week of May or the

beginning of June. Warmer May temperatures may result in earlier melt-out, rapid spring

overturning, and an extended spring growing season. Additionally, earlier melt-out may

result in increased overland-flow and near-surface through-flow entering the lake early in

the season. If the ground is still frozen and lacks vegetation cover, precipitation may take

a more direct, near-surface route to the lake.

In southwest Alaska, spring temperature and regional precipitation are directly

influenced by the AL, dominant feature of winter atmospheric and oceanic circulation in

the north Pacific Ocean. A strengthened AL tends to be associated with warmer and

wetter winters along the Bering Sea due to the increased tendency for storm systems to

pump warm air poleward (Rodionov et al., 2005). Additionally, of the climate modes

defined by Mock et al. (1998), the northeast Pacific negative type and the northcentral

Pacific negative type are the only two that result in positive winter temperature and

precipitation anomalies in southwestern Alaska, and these two modes feature a

strengthened AL.

Page 59: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

51

The annual NPI is also a measure of AL strength, yet it shows no correlation with

the BSi content of Cascade Lake (r = 0.17, p = 0.229). Although the ALPI and NPI are

both measures of AL strength, they are negatively correlated (r = -0.48, p = 3.46 x 10-7)

because they relate to AL variations differently. The ALPI is a measure of the average

ocean area affected by low pressures (<100.5 kPa) from December-March, whereas the

annual NPI is the area-weighted sea-level pressure (Beamish, 1997; Tenberth and

Hurrell, 1994). Because the NPI is based on annual statistics, not only winter, the index is

also influenced by the ENSO cycles of the South Pacific and therefore is less directly

related to the climate regimes that affect algal growth at Cascade Lake. However, if the

NPI index is calculated over the same months as the ALPI (DJFM), then the two indices

are more strongly correlated (r = -0.73, p = 1.41 x 10-18) and the correlation between BSi

flux and the DJFM NPI five-year running average becomes significant (r = -0.42, p =

0.002). The ALPI shows moderately strong correlations with the Dillingham winter

precipitation and May temperatures (r = 0.33, p = 0.056, and r = 0.31, p = 0.011,

respectively).

Paleoclimate Interpretations. The AL is likely the dominant factor in the regional climate

that affects BSi at Cascade Lake, both in the last century and throughout the late

Holocene. If so, then BSi fluctuations in the sediment cores can be used to characterize

variations in the AL over the late Holocene with high BSi values being interpreted as

‘mild’ climates (moist and warm, with a strengthen AL). This interpretation of BSi flux

was completed at a sub-decadal resolution back to ~2000 cal yr BP based on the high-

resolution sampling interval (Figure 27). BSi flux shows three major fluctuations over

Page 60: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

52

this period. Lake productivity was generally low from 2000 to 1450 cal yr BP (BSi flux

= 1.4 mg·cm-2·yr-1) and rose from 1450 to 1150 cal yr BP (2.6 mg·cm-2

·yr-1). BSi flux

decreased from 1150 cal yr BP to 1850 AD (1.9 mg·cm-2·yr-1) and subsequently increased

to its highest value in 1995 (4.4 mg·cm-2·yr-1). The pattern of these fluctuations (not the

absolute timing) corresponds to the general climate patterns of the late Holocene,

primarily the Medieval Warm Period (MWP), the Little Ice Age, and 20th century

warming. Additionally, the climate patterns of the LIA interpreted from BSi flux are

supported by the geomorphic evidence for glacial fluctuations. Low BSi flux in the

1800’s corresponds to ice advance in the study area, followed by subsequent ice retreat

and moraine stabilization in ~1860, and a return to increased BSi flux (Figure 27). The

high productivity during the mid-1990’s in Cascade Lake was unequalled at any time in

the last 2000 yr and may indicate that the modern strengthening of the AL is significantly

greater than during the past.

Few high-resolution records of climate continuously span the last few millennia in

Alaska. The most complete and comparable records are that of Hu et al. (2001) and Loso

et al. (2006) in southern Alaska (Figure 2), and Anderson et al. (2005) in the southern

Yukon territory. Hu et al.’s (2001) record of oxygen isotopes from Farewell Lake

(western Alaska Range) indicates a relatively warm climate from 850 to 1200 AD (1150-

750 cal yr BP). This agrees with the general varve thickening in Iceberg Lake (Loso et

al., 2006) from 1000 to 1250 AD (1000-800 cal yr BP) and the AL strengthening

interpreted at Jellybean Lake (Anderson et al., 2005) between 1200 and 900 cal yr BP.

This warm period is also indicated by the general lack of glacial advances across the

Northern Hemisphere (Porter, 1986). If the rise in BSi flux between 1450 and 1150 cal

Page 61: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

53

yr BP in Cascade Lake is related to the mild climate of the MWP, then the timing of

MWP initiation is not synchronous and shows progressively later initiation eastward

across southern Alaska over a 450 yr period. Cascade Lake is ~900 km west of Iceberg

Lake and 600 km southwest of Farewell Lake (Figure 2). Warming at Cascade Lake was

initiated 200 yr prior to that at Farewell Lake, 400 yr prior to that of Iceberg Lake, and

250 yr prior to Jellybean Lake (Figure 28).

The time transgression of MWP initiation may be explained by three possible

causes, or some combination thereof. (1) Each of the climate records discussed above

derive from different proxy records (varve thickness in Iceberg Lake, isotopes in

Farewell Lake and Jellybean Lake, and BSi in Cascade Lake). While each of these

proxies record climate variations, they may each respond to different combinations of

temperature and precipitation. Depending on the climate thresholds necessary to produce

a measurable change, each proxy may record the initiation of different ‘stages’ of the

climate event. (2) The age model for Cascade Lake over this time period is based on a

linear extrapolation of sedimentation rates to the Aniakchak tephra deposited at 3500 cal

yr BP. Sedimentation rates likely vary with varying climate regimes, and the assumption

of a linear sedimentation rate over this extended period inserts some uncertainty in the

Cascade Lake age model. (3) The average position of the AL migrated westward during

the MWP.

Extending the climate interpretations beyond the last 2000 yr in Cascade Lake

suggests other general late Holocene climate trends (Figure 28). BSi flux prior to ~6000

cal yr BP is low and shows little variation. This may, reflect, the flux calculation using

the low sedimentation rate below Tephra B, or might indicate limited nutrient input to the

Page 62: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

54

lake. Following deglaciation, the lack of vegetation (specifically the lack of nitrogen-

fixing alder) would limit nutrient levels in the drainage basin. OM values increase

between 11,000 and 7000 cal yr BP, indicating a milder climate following deglaciation

and the cool period between 13,000-12,000 cal yr BP that has been inferred from lowered

productivity in Arolik Lake during that time (Kaufman et al., 2001). Both Waskey Lake

(Levy et al., 2004) and Cascade Lake record the highest OM preservation between 6500

and 7500 cal yr BP. Waskey Lake also experience limited glacial input during this time,

making this period a likely candidate for the regional Holocene thermal maximum. This

is significantly younger than the average 10,000 cal yr BP age of the Holocene thermal

maximum across central and eastern Beringia (Kaufman et al., 2004).

Both Waskey Lake and Cascade Lake also record cooler climates between 3000

and 2500 cal yr BP, after the deposition of the Aniakchak tephra, marking the regional

initiation of Neoglaciation. This age is younger than initial Neoglacial advances

elsewhere in Alaska (4000-3500 cal yr BP in the Brooks Range, Ellis and Calkin, 1984;

4000-3000 cal yr BP on the Seward Peninsula, Calkin et al., 1998) and indicates a time-

transgressive climate shift potentially similar to that of the Medieval Warm Period.

Following the initiation of Neoglaciation, the glacial records at Waskey Lake and

Cascade Lake areas diverge, with multiple Neoglacial moraines in the headwaters of

Waskey Lake and only a single moraine formed late during the LIA around Cascade

Lake. As discussed previously, the cause of this discrepancy is uncertain, though it may

be related to variations in glacier flow mechanics caused by different subglacial bedrock

lithologies.

Page 63: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

55

CHAPTER 6:

CONCLUSIONS

This study combines a moraine record of glacial fluctuations with an

interpretation of the sediments of Cascade Lake to determine the variability of Holocene

climate in the Ahklun Mountains. The glaciers of the Cascade Lake area reached their

maximum Neoglacial extent during the LIA. Lichenometric ages of five moraines

deposited by glaciers draining into Cascade Lake were derived using the Seward

Peninsula growth curve adapted from Solomina and Calkin (2003) with an increased

colonization time (~30 years). These ages place the stabilization of maximum LIA

moraines at ~1860 AD. Following moraine stabilization, Cascade Glacier retreated at a

variable rate, averaging 11 m/yr, to its current location 1.4 km up-valley.

During the LIA, cirque and valley glaciers of the Cascade Lake area experienced

an average ELA lowering of ~22 ± 12 m (n = 8) relative to 1957 AD. The ∆ELA values

of this study agree with other studies (Levy et al., 2004; Skiorski, 2004; Daigle and

Kaufman, 2006) that used similar methodologies to derive the ‘modern’ ELA. This

minimal ∆ELA in the Cascade Lake area probably reflects limited winter precipitation

during the LIA in southwestern Alaska.

The study of modern sedimentation patterns and temperatures of Cascade Lake

provides a context for interpreting late Holocene lake sediment. The contrast in sediment

accumulation and BSi flux between the two collection years (70% and 77% less,

respectively, in the cooler 2006 spring/summer) indicates the strong control of

spring/summer conditions on sediment and BSi accumulation in the lake. No

precipitation data are available for Cascade Lake or Dillingham, making it difficult to

Page 64: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

56

determine whether this difference is related to temperature or to precipitation.

Precipitation has a more direct impact on discharge and terrestrial input of nutrients to the

lake, but both of these factors likely affect sediment and BSi accumulation to varying

degrees.

Annual fluctuations in the BSi, OM, and grain size analyzed in the trapped 2004-

2006 sediments indicate a seasonal signal is preserved in the sediments. Sediment with

low (high) BSi values correspond to elevated (decreased) bulk densities, and decreased

(increased) OM and grain size, and was probably deposited during the fall/winter

(spring/summer).

Of the three sediment cores recovered from Cascade Lake, I focused on Core 02

due to its longer record (compared to Core 03) and the occurrence of laminations. 210Pb,

137Cs, and the provisional identifications of the Aniakchak tephra and Tephra B provide

the age control for the core. The chronology below Tephra B is extrapolated using a

linear sedimentation rate. Overall, the chronologic control is limited and uncertain. MS

and OM values show relatively little variation down core compared to the flux of BSi,

and laminations are multi-annular (not varved). Therefore, I focused on BSi flux for the

purpose of paleoclimate reconstruction due to its apparent sensitivity to climate

variability and the ability to measure BSi at a high resolution (sub-decadal).

BSi flux (avg. 3 yr sampling resolution) was analyzed over the last 150 yr and

correlated with a variety of instrumental climate parameters. The climatological

parameters most strongly correlated with algal production in Cascade Lake are river

discharge at a nearby gaging station, spring temperature, and the strength of the AL, with

algal production favored during moist and warm conditions (strengthened AL). High BSi

Page 65: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

57

flux indicates a mild climate from 1450-1150 cal yr BP (MWP?), a cool climate from

1150 cal yr BP to 1850 AD (LIA?), and a 20th century climate that was milder than that

of even the MWP.

If the interpretation of climate variation is extended to the full Holocene sequence

of lake sediments, other long-term trends can be inferred. The core did not extend into

ice-proximal sediments, so conditions immediately following deglaciation are not known.

Eventually the deglacial climate warmed and reached a period of maximum warmth

~7000 cal yr BP. Neoglacial cooling was initiated ~3000 cal yr BP, and glaciers reached

their maximum LIA extent during the late 19th century.

Interpretations of sediments from Cascade Lake and the glaciers in the

surrounding area indicate several periods of climate variability in the Ahklun Mountains

during the late Holocene. This variability is closely related to the dynamics of the North

Pacific Ocean and provides an important context for interpreting current and potential

future climate change. No other interval of the Cascade Lake sediment record is

comparable to the most recent period of warm and wet conditions, suggesting an

unprecedented strengthening of the AL during the 20th century.

Page 66: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

58

References Cited

Abbott, M.B., Finney, B.P., Edwards, M.E., Kelts, K.R., 2000. Lake-level reconstructions and paleohydrology of Birch Lake, central Alaska, based on seismic reflection profiles and core transects. Quaternary Research 53, 154-166.

Alley, R.B., Mayewski, P.A., Sowers, T., Stuiver, M., Taylor, K.C., Clark, P.U., 1997.

Holocene climatic instability: A prominent, widespread event 8200 yr ago. Geology 25, 483-486.

Anderson, N.J., 2000. Diatoms, temperature and climate change. European Journal of

Phycology 35, 307-314. Anderson, L., Abbott, M.B., Finney, B.P., Burns, S.J., 2005. Regional atmospheric

circulation change in the North Pacific during the Holocene inferred from lacustrine carbonate oxygen isotopes, Yukon Territory, Canda. Quaternary Research 64, 21-35.

Appleby, P.G., 2004. Chronostratigraphic techniques in recent sediments, in Last, W.M.,

Smol, J.P. (eds), Tracking Environmental Change Using Lake sediments: Volume 1. Basin Analysis, Coring and Chronological Techniques, 171-203 (Dordrecht: Kluwer).

Axford, Y., Kaufman, D.S., 2004. Late glacial and Holocene glacier and vegetation

fluctuations at Little Swift Lake, southwestern Alaska, U.S.A. Arctic, Antarctic, and Alpine Research 36, 139-146.

Bakke, J., Lie, O., Nesje, A., Dahl, S.O., Paasche, O., 2005. Utilizing physical sediment

variability in glacier-fed lakes for continuous glacier reconstructions during the Holocene, northern Folgefonna, western Norway. The Holocene 15, 161-176.

Béget, J., Mason, O., Anderson, P., 1992. Age, extent and climatic significance of the c.

3400 BP Aniakchak tephra, western Alaska, USA. The Holocene 2, 51-56. Beamish, R.J., Neville, C.E., Cass, A.J., 1997. Production of Fraser River sockeye

salmon (Oncorhynchus nerka) in relation to decadal-scale changes in the climate and the ocean. Canadian Journal of Fisheries and Aquatic Science 54, 543-554. Digital data at: http://www.pac.dfo-mpo.gc.ca/sci/sa-mfpd/downloads/indices/alpi.txt.

Bianchi, G.G., McCave, I.N., 1999. Holocene periodicity in North Atlantic climate and

deep-ocean flow south of Iceland. Nature 397, 515-517.

Page 67: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

59

Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas, I., Bonani, G., 2001. Persistent solar influence on North Atlantic climate during the Holocene. Science 294, 2130-2136.

Bond, G., Showers, W., 1997. A pervasive millennial-scale cycle in North Atlantic

Holocene and glacial climates. Science 278, 1257-1266. Borchardt, G.A., Aruscavage P.J., Millard, H.T., 1972. Correlation of the Bishop ash, a

Pleistocene marker bed, using instrumental neutron activation analysis. Journal of Sedimentary Petrology 42, 301-306.

Box, S.E., Moll-Stalcup, E.J., Frost, T.P., Murphy, J.M., 1993. Preliminary geologic map

of the Bethel and southern Russian Mission quadrangles, Southwestern Alaska. U.S. Geologic Survey Report: MF-2226-A, 20 pp.

Bradbury, P., Cumming, B., Laird, K., 2002. A 1500-year record of climatic and

environmental change in Elk Lake, Minnesota III: measures of past primary productivity. Journal of Paleolimnology 27, 321-340.

Briner, J.P., Kaufman, D.S., Werner, A., Caffee, M., Levy, L., Kaplan, M.R., and Finkel,

R.C., 2002. Glacier readvance during the late glacial (Younger Dryas?) in the Ahklun Mountains, southwestern Alaska. Geology 30, 679-682.

Briner, J.P., Kaufman, D.S., 2000. Late Pleistocene glaciation of the southwestern

Ahklun Mountains, Alaska. Quaternary Research 53, 13-22. Calkin, P.E., Wiles, G.C., Barclay, D.J., 2001. Holocene coastal glaciation of Alaska.

Quaternary Science Reviews 20, 449-461. Calkin, P.E., 1988. Holocene glaciation of Alaska (and adjoining Yukon Territory,

Canada). Quaternary Science Reviews 7, 159-184. Calkin, P.E., Haworth, L.A., Burns, P.E., 1985. Cirque glacier regime and Neoglaciation,

Brooks Range, Alaska. Zeitschrift fur Gletscherkude und Glazialgeologie 21, 371-378.

Calkin, P.E., Ellis, J.M., 1980. A lichenometric dating curve and its application to

Holocene glacier studies in the central Brooks Range, Alaska. Arctic and Alpine Research 12, 245-264.

Campbell, I.D., Campbell, C., Apps, M.J., Rutter, N.W., Bush, A.B.G., 1998. Late

Holocene ~1500 yr climatic periodicities and their implications. Geology 26, 471-473.

Colman, S.M., Bratton, J.F., 2003. Anthropogenically induced changes in sediment and

biogenic silica fluxes in Chesapeake Bay. Geology 31, 71-74.

Page 68: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

60

Crowley, T.J., Lowery, T.S., 2000. How warm was the Medieval Warm Period? Ambio

29, 51-54. Dahl, S.O., Nesje, A., 1992. Paleoclimatic implications based on equilibrium-line

altitude depressions of reconstructed Younger Dryas and Holocene cirque glaciers in inner Nordfjord, western Norway. Palaeogeography, Palaeoclimatology, Palaeoecology 94, 87-97.

Daigle, T.A., Kaufman, D.S., 2007. Late Holocene climate change at Goat Lake, Kenai

Mountains, south-central Alaska. American Geophysical Union, Programme with Abstracts, in press.

D’Arrigo, R., Wilson, R., Deser, C., Wiles, G., Cook, E., Villalba, R., Tudhope, A., Cole,

J., Linsley, B., 2005. Tropical-North Pacific climate linkages over the past four centuries. Journal of Climate 18, 5253-5265.

Davi, N.K., Jacoby, G.C., Wiles, G.C., 2003. Boreal temperature variability inferred

from maximum latewood density and tree-ring width data, Wrangell Mountain region, Alaska. Quaternary Research 60, 252-262.

Dean, W.E., 1974. Determination of carbonate and organic matter in calcareous

sediments and sedimentary rocks by loss on ignition: comparison with other methods. Journal of Sedimentary Petrology 44, 242-248.

Dearing, J., 1999. Environmental magnetic susceptibility: Using the Bartington MS2

system. Chi Publishing, Kenilworth, England. Denton, G.H., Karlén, W., 1973. Holocene climatic variations—Their pattern and

possible cause. Quaternary Research 3, 155-205. Douglas, M.S.V., Hamilton, P.B., Pienitz, R., Smol, J.P., 2004. Algal indicators of

environmental change in arctic and Antarctic lakes and ponds, in Pienitz, R. Douglas, M.S.V., Smol, J.P. (eds), Long-term Environmental Change in Arctic and Antarctic Lakes, 117-158 (Springer: Dordrecht)

Downing, J.A., Rath, L.C., 1988. Spatial patchiness in the lacustrine sedimentary

environment. Limnology and Oceanography 33, 447-458. Ebbesmeyer, C.C., Strickland, R.M., 1995. Oyster condition and climate: Evidence from

Willapa Bay. Publication WSG-MR 95-02, Washington Sea Grant Program, University of Washington, 11 p. Digital data at: www.cqs.washington.edu/data.

Ellis, J.M., Calkin, P.E., 1984. Chronology of Holocene glaciation, central Brooks

Range, Alaska. Geological Society of America Bulletin 95, 897-912.

Page 69: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

61

Engstrom, D.R., Fritz, S.C., 2006. Coupling between primary terrestrial succession and the tropic development of lakes at Glacier Bay, Alaska. Journal of Paleolimnology 35, 873-880.

Evison, L.H., Calkin, P.E., Ellis, J.M., 1996. Late-Holocene glaciation and twentieth-

century retreat, northeastern Brooks Range, Alaska. The Holocene 6, 17-24. Francus, P., Bradley, R.S., Abbott, M.B., Patridge, W., Keimig, F., 2002. Paleoclimate

studies of minerogenic sediments using annually resolved textural parameters. Geophysical Research Letters 29, 1-4.

Grove, J.M., 2003. Little Ice Ages: Ancient and Modern, pp. 643 (Routledge: Taylor &

Francis Group). Hardy, D.R., Bradley, R.S., Zolitschka, B., 1996. The climatic signal in varved

sediments from Lake C2, northern Ellesmere Island, Canada. Journal of Paleolimnology 16, 227-238.

Hansen, B.C.S, Engstrom, D.R., 1996. Vegetation history of Pleasant Island,

southeastern Alaska, since 13,000 yr B.P. Quaternary Research 46, 161-175. Hoare, J.M., Coonrad, W.L., 1959. Geology of the Bethel quadrangle, Alaska. U.S.

Geologic Survey Miscellaneous Geological Investigations Map I-285. Hodgkins, R., Cooper, R., Wadham, J., Tranter, M., 2003. Suspended sediment fluxes in

a high-Arctic glacierised catchment: implications for fluvial sediment storage. Sedimentary Geology 162, 105-117.

Hu, F.S., Finney, B.P., Bubaker, L.B., 2001. Effects of Holocene Alnus expansion on

aquatic productivity, nitrogen cycling, and soil development in southwestern Alaska. Ecosystems 4, 358-368.

Hu, F.S., Ito, E., Brown, T.A., Curry, B.B., Engstrom, D.R., 2001. Pronounced climatic

variations in Alaska during the last two millennia. Proceedings of the National Academy of Sciences 19, 10552-10556.

Hu, F.S., Kaufman, D., Yoneji, S., Nelson, D., Shemesh, A., Huang, Y., Tian, J., Bond,

G., Clegg, B., Brown, T., 2003. Cyclic variation and solar forcing of Holocene climate in the Alaska subarctic. Science 301, 1890-1894.

Hu, F.S., Shemesh, A., 2003. A biogenic silica δ18O record of climatic change during the

last glacial-interglacial transition in southwestern Alaska. Quaternary Research 59, 379-385.

Page 70: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

62

Hu, F.S., Nelson, D.M., Clarke, G.H., Rühland, K.M., Huang Y., Kaufman, D.S., Smol, J.P., 2006. Abrupt climatic events during the last glacial-interglacial transition in Alaska. Geophysical Research Letters 33, L18708.

Hughen, K.A., Overpeck, J.T., Anderson, R.F., 2000. Recent warming in a 500-year

palaeotemperature record from varved sediments, Upper Soper Lake, Baffin Island, Canada. The Holocene 10, 9-19.

Innes, J.L., 1985. Lichenometry. Progress in Physical Geography 9, 187-254. Kaufman, D.S., Ager, T.A., Anderson, N.J., Anderson, P.M., Andrews, J.T., Bartlein,

P.J., Brubaker, L.B., Coats, L.L., Cwynar, L.C., Duvall, M.L., Dyke, A.S., Edwards, M.E., Eisner, W.R., Gajewski, K., Geirsdottir, A., Hu, F.S., Jennings, A.E., Kaplan, M.R., Kerwin, M.W., Lozhkin A.V., MacDonald, G.M., Miller, G.H., Mock, C.J., Oswald, W.W., Otto-Bliesner, B.L., Porinchu, D.F., Ruhland, K., Smol, J.P., Steig, E.J., Wolfe, B.B., 2004. Holocene thermal maximum in the western Artic (0-180°W). Quaternary Science Review 23, 529-560.

Kaufman, D.S., Hu, F.S., Briner, J.P., Werner, A., Finney, B.P., and Gregory-Eave, I.,

2003. A ~33,000 year record of environmental change from Arolik Lake, Ahklun Mountains, Alaska. Journal of Paleolimnology 30, 343-362.

Kaufman , D.S., Manley, W.F., Forman, S.L., Layer, P.W., 2001. Pre-Late-Wisconsin

glacial history, coastal Ahklun Mountains, southwestern Alaska-new amino acid, thermoluminescence, and 40Ar/39Ar results. Quaternary Science Reviews 20, 337-352.

Kayano, M.T., Brahmananda Rao, V., Andreoli, R.V., 2005. A review of short-term climate variability mechanisms. Advances in Space Research 35, 843-851.

Kim, B.Y., Kodama, K.P., Moeller, R.E., 2005. Bacterial magnetite produced in water

column dominates lake sediment mineral magnetism: Lake Ely, USA. Geophysical Journal International 163, 26-37.

Lamoureux, S.F., 1999. Catchment and lake controls over the formation of varves in

monomictic Nicolay Lake, Cornwall Island, Nunavut. Canadian Journal of Earth Science 36, 1533-1546.

Lamoureux, S.F., Gilbert, R., 2004. Physical and chemical properties and proxies of high

latitude lake sediments, in Long-term Environmental Change in Arctic and Antarctic Lakes, 53-88 (Springer: Dordrecht).

Lafrenière, M.J., Sharp, M.J., 2005. A comparison of solute fluxes and sources from

glacial and non-glacial catchments over contrasting melt seasons. Hydrological Processes 19, 2991-3012.

Page 71: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

63

Leemann, A., Niessen, F., 1994. Varve formation and the climatic record in an alpine proglacial lake: calibrating annually-laminated sediments against hydrological and meteorological data. The Holocene 4, 1-8.

Lemmen, D.S., Gilbert, R., Smol, J.P., Hall, R.I., 1998. Holocene sedimentation in

glacial Tasikutaaq Lake, Baffin Island. Canadian Journal of Earth Science 25, 810-823.

Leonard, E.M., Reasoner, M.A., 1999. A continuous Holocene glacial record inferred

from proglacial lake sediments in Banff National Park, Alberta, Canada. Quaternary Research 51, 1-13.

Levy, L.B., Kaufman, D.S., Werner, A., 2004. Holocene glacier fluctuations, Waskey

Lake, northeastern Ahklun Mountains, southwestern Alaska. The Holocene 14, 185-193.

Levy, L.B., 2002. Late Holocene glacier fluctuations, Northeastern Ahklun Mountains,

southwestern Alaska. Master’s thesis, Northern Arizona University. Locke, W.W. II., Andrews, J.T., Webber, P.J., 1979. A manual for lichenometry. British

Geomorphological Research Group Technical Bulletin 26, 47 p. Loso, M.G., Anderson, R.S., Anderson, S.P., Reimer, P.J., 2006. A 1500-year record of

temperature and glacial response inferred from varved Iceberg Lake, southcentral Alaska. Quaternary Research 66, 12-24.

Lotter, A.F., Pienitz, R., Schmidt, R., 1999. Diatoms as indicators of environmental

change near arctic and alpine treeline, in Stoermer, E.F., Smol, J.P. (eds), The Diatoms: Applications for the Environmental and Earth Sciences, 205-226 (Cambridge: Cambridge University Press).

Maasch, K.A., Mayewski, P.A. Rohling, E.J., Stager, J.C., Karlén, W., Meeker, L.D.,

Meyerson, E.A., 2005. A 2000-year context for modern climate change. Geografiska Annaler 87, 7-15.

Manley, W.F., Kaufman, D.S., Briner, J.P., 2001. Pleistocene glacial history of the

southern Ahklun Mountains, southwestern Alaska: Soil-development, morphometric, and radiocarbon constraints. Quaternary Science Reviews 20, 353-370.

Manley, W.F., 1999. GIS assessment of glaciers, equilibrium-line altitudes, and climate

sensitivity, an example from southwestern Alaska. Geological Society of America, Abstracts with Programs 31, 366.

Mantua, N.J., Hare, S.R., Zhang, Y., Wallace, J.M., Francis, R.C., 1997. A Pacific

interdecadal climate oscillation with impacts on salmon production. Bulletin of

Page 72: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

64

the American Meteorological Society 78, 1069-1079. Digital data at: http://jisao.washington.edu/pdo/PDO.latest.

Mayewski, P.A., Rohling, E.E., Stager, J.C., Karlen, W., Maasch, K.A., Meeker, L.D.,

Meyerson, E.A., Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F., Staubwasser, M., Schneider, R.R., Steig, E.J., 2004. Holocene climate variability. Quaternary Research 62, 243-255.

Mayewski, P.A., Meeker, L.D., Twickler, M.S., Whitlow, S., Yang, Q., Lyons, W.B.,

Prentice, M., 1997. Major features and forcing of high-latitude northern hemisphere atmospheric circulation using a 110,000-year long glaciochemical series. Journal of Geophysical Research 102, 2634-26366.

Meier, M.F., Post, A.S., 1962. Recent variations in mass net budgets of glaciers in

western North America. International Association of Scientific Hydrology 58, 63-77.

Minobe, S., Mantua, N., 1999. Interdecadal modulation of interannual atmospheric and

oceanic variability over the North Pacific. Progress in Oceanography 43, 163-192.

Mock, C.J., Bartlein, P.J., Anderson, P.M., 1998. Atmospheric circulation patterns and

spatial climatic variations in Beringia. International Journal f Climatology 10, 1085-1104.

Moore, J.J., Hughen, K.A., Miller, G.H., Overpeck, J.T., 2001. Little Ice Age recorded in

summer temperature reconstruction from varved sediments of Donard Lake, Baffin Island, Canada. Journal of Paleolimnology 25, 503-517.

Mortlock, R.A., Froelich, P.N., 1989. A simple method for the rapid determination of

biogenic opal in pelagic marine sediments. Deep-Sea Research 36, 1415-1426. NCAR, 2006. NP data provided by the Climate Analysis Section, NCAR, Boulder, USA,

Trenberth and Hurrell (1994). Digital data at: www.cgd.ucar.edu/cas/jhurrell/npindex.html.

NCDC, 2005. National climatic data center data set for daily surface data, Dillingham,

AK. Digital data at: http://www.ncdc.noaa.gov/oa/ncdc.html. Neal, E.G., Walter, M.T., Coffeen, C., 2002. Linking the pacific decadal oscillation to

seasonal stream discharge patterns in southeast Alaska. Journal of Hydrology 263, 188-197.

Noon, P.E., Birks, H.J.B., Jones, V.J., Ellis-Evans, J.C., 2001. Quantitative models for

reconstructing catchment ice-extent using physical-chemical characteristics of lake sediments. Journal of Paleolimnology 25, 375-392.

Page 73: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

65

Noren, A.J., Bierman, P.R., Steig, E.J., Lini, A., Southon, J.A., 2002. Millennial-scale

storminess variability in the northeastern United States during the Holocene. Nature 419, 821-824.

NWIS, 2006. National Water Information System data set for Nuyakuk River discharge

1949-2004. Digital data at: http://waterdata.usgs.gov/ak/nwis/. O’Brien, S.R., Mayewski, P.A., Meeker, L.D., Meese, D.A., Twickler, M.S., Whitlow,

S.I., 1995. Complexity of Holocene climate as reconstructed from a Greenland ice core. Science 270, 1962-1964.

O’Neal, M.A., 2006. The effects of slope degradation on lichenometric dating of Little

Ice Age moraines. Quaternary Geochronology 1, 121-128. Overpeck, J., Hughen, K., Hardy, D., Bradley, R., Case, R., Douglas, M., Finney, B.,

Gajewski, K., Jacoby, G., Jennings, A., Lamoureux, S., Lasca, A., MacDonald, G., Moore, J., Retelle, M., Smith, S., Wolfe, A., Zielinski, G., 1997. Arctic environment change of the last four centuries. Science 278, 1251-1257.

Papineau, J.M., 2001. Wintertime temperature anomalies in Alaska orrelated with ENSO

and PDO. International Journal of Climatology 21, 1577-1592. PNWPEST, 2006. Current and historical weather data set 2004-2006, maximum and

minimum daily temperature for Dillingham, Alaska and Bethel, Alaska. Integrated Plant Protection Center, Oregon State University. Digital data set: http://pnwpest.org/ak/.

Porter, S.C., 2001. Snowline depression in the tropics during the last glaciation.

Quaternary Science Reviews 20, 1067-1091. Porter, S.C., 1986. Pattern and forcing of Northern Hemisphere glacier variations during

the last millennium. Quaternary Research 26, 27-48. Porter, S.C., Denton, G.H., 1967. Chronology of Neoglaciation in the North American

cordillera. American Journal of Science 255, 177-210. Pienitz, R., Douglas, M.S.V., Smol, J.P., eds., 2004. Long-term Environmental Change in

Arctic and Antarctic Lakes, v. 8. Springer, Dorderecht, 509 p. Qiu, L., Williams, D.F., Gvorzdkov, A., Karabanov, E., Shimaraeva, M., 1993. Biogenic

silica accumulation and paleoproductivity in the northern basin of Lake Baikal during the Holocene. Geology 21, 25-28.

Page 74: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

66

Riehle, J.R., Meyer, C.E., Ager, T.A., Kaufman, D.S., Ackerman, R.E., 1987. The Aniakchak tephra deposit, a late Holocene marker horizon in western Alaska. US Geological Survey Circular 998, 9-22.

Rodionov, S.N., Overland, J.E., Bond, N.A., 2005. The Aleutian low and winter climatic

conditions in the Bering Sea. Part I: Classification. Journal of Climate 18, 160-177.

Ryves, D.B., Juggins, S., Fritz, S.C., Battarbee, R.W., 2001. Experimental diatom

dissolution and the quantification of microfossil preservation in sediments. Palaeogeography, Palaeoclimatology, Palaeoecology 172, 99-113.

SCAS, 2005. PRISM data set 1961-1990, mean monthly precipitation and temperature

for Alaska and Yukon Territory, Spatial Climate Analysis Service, Oregon State University. Digital data set: http://www.oc.orst.edu/prism/.

Schneider, N., Cornuelle, B.D., 2006. The forcing of the Pacific Decadal Oscillation.

Journal of Climate 18, 4355-4373. Skiorski, J.J., 2004. Little Ice Age glacier fluctuations and winter precipitation, Brooks

Range, Alaska. Master of Science thesis, Northern Arizona University. Smol, J.P., Cumming, B.F., 2000. Tracking long-term changes in climate using algal

indicators in lake sediments. Journal of Phycology 36, 986-1011. Solomina, O., Calkin, P.E., 2003. Lichenometry as applied to moraines in Alaska,

U.S.A., and Kamchatka, Russia. Arctic, Antarctic and Alpine Research 35, 129-143.

Spooner, I.S., Barnes, S., Baltzer, K.B., Raeside, R., Osborn, G.D., Mazzucchi, 2003.

The impact of air mass circulation dynamics on late Holocene paleoclimate in northwestern North America. Quaternary International 108, 77-83.

Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A. et al., 1998.

INTCAL98 Radiocarbon age calibration, 24,000-0 cal BP. Radiocarbon 40, 1041-1082.

Swann, G.E.A., Mackay, A.W., 2006. Potential limitations of biogenic silica as an

indicator of abrupt climate change in Lake Baikal, Russia. Journal of Paleolimnology 36, 81-89.

Telford, R.J., Barker, P., Metcalfe, S., Newton, A., 2004. Lacustrine responses to tephra

deposition: examples from Mexico. Quaternary Science Reviews 23, 2337-2353. Trenberth, K.E., Hurrell, J.W., 1994. Decadal atmosphere-ocean variation in the Pacific.

Climate Dynamics 9, 303-319.

Page 75: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

67

Torsnes, I., Rye, N., Nesje, A., 1993. Modern and Little Ice Age equilibrium-line

altitudes on outlet valley glaciers from Jostedalsbreen, Western Norway: an evaluation of different approaches to their calculation. Arctic and Alpine Research 25, 106-116.

Verschuren, D., 1999. Sedimentation controls on the preservation and time resolution of

climate-proxy records from shallow fluctuating lakes. Quaternary Science Reviews 18, 821-837.

Waythomas, C.F., Neal, C.A., 1998. Tsunami generation by pyroclastic flow during the

3500-year B.P. caldera-forming eruption of Aniakchak Volcano, Alaska. Bulletin of Volcanology 60, 110-124

Werner, A., 1990. Lichen growth rates for the northwest coast of Spitsbergen, Svalbard.

Arctic and Alpine Research 22, 129-140. Wiles, G.C., Calkin, P.E., 1994. Late Holocene, high-resolution glacial chronologies and

climate, Kenai Mountains, Alaska. Geological Society of America Bulletin 106, 281-303.

Willemse, N.W., Tornqvist, T.E., 1999. Holocene century-scale temperature variability

from West Greenland lake records. Geology 27, 580-584. WRCC, 2006. Western Regional Climate Center data set of historical climate data for

Dillingham, AK FAA airport. Digital data at: http://www.wrcc.dri.edu/index.html.

Page 76: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

68

TABLE 1. INSTRUMENTS DEPLOYED AT CASCADE LAKE Instrument Location Latitude

Longitude (NAD 27)

Launch date

Recovery date

Comments

Mooring 1 Sediment trap Hobo Water

Temp Pro

Core site 01

~3 m water depth

Sediment trap

Hobo Water Temp Pro

Core site

01 ~23 m

water depth

60.14231 159.41209

6/28/04

8/12/05

8/10/05

7/19/06

Sediment accumulation =0.11 g/cm3/yr in

2004-2005

Mooring 2 Sediment trap Hobo Water

Temp Pro

Core site 03

~3 m water depth

Sediment trap Hobo Water

Temp Pro

Core site

03 ~51 m

water depth

60.13643 159.39917

6/28/04

8/12/05

8/11/05

7/19/06

Sediment accumulation =0.10 g/cm3/yr in

2004-2005 = 0.03 g/cm3/yr in

2005-2006

Air temperature Hobo Pro Western

lake shore (Fig. 3)

60.14082 159.41236

06/28/4

8/12/05

8/9/05

7/20/06

Temp. logger failed between 12/22/04 and 8/9/05

Hobo Pro Cascade Glacier ice proximal (Fig. 3)

60.08872 159.34043

8/13/05 7/20/06

Page 77: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

69

TABLE 2. PHYSICAL AND BIOLOGICAL DATA FROM BOTTOM SEDIMENT TRAPS Unit Top

depth (cm)

Bottom depth (cm)

BSi

(%)

BSi flux

(mg·cm-2·yr-1)

Median grain size

(µm)

Organic matter (%)

Bulk density

(g/cm3) Mooring 1: 2004-2005 1 0.0 1.9 2.91 3.20 5.19 5.5 0.29 2 1.9 3.9 2.50 2.75 1.95 6.1 0.18 3 3.9 7.9 1.86 2.05 3.62 4.5 0.34 4 7.9 9.4 1.49 1.64 4.55 3.5 0.86 5 9.4 14.4 2.04 2.25 4.61 4.0 0.60 Mooring 2: 2004-2005 1 0.0 1.7 N.D. N.D. N.D. N.D. 0.40 2 1.7 8.9 N.D. N.D. N.D. N.D. 0.25 3 8.9 11.9 N.D. N.D. N.D. N.D. 0.44 4 11.9 12.9 N.D. N.D. N.D. N.D. 0.47 5 12.9 15.9 N.D. N.D. N.D. N.D. 0.31 6 15.9 17.7 N.D. N.D. N.D. N.D. 0.48 7 17.7 18.1 N.D. N.D. N.D. N.D. 0.66 Mooring 2: 2005-2006 1 0.0 1.6 1.27 0.38 N.D. N.D. 0.22 2 1.6 2.3 1.60 0.48 N.D. N.D. 0.37 3 2.3 2.5 1.59 0.48 N.D. N.D. 0.56 4 2.5 5.9 1.67 0.50 N.D. N.D. 0.11 5 5.9 8.1 1.30 0.39 N.D. N.D. 0.18 6 8.1 8.5 1.60 0.48 N.D. 0.66 Note: N.D. = no data collected on these samples.

Page 78: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

70

TABLE 3. CLIMATE DATA FROM PREVIOUSLY PUBLISHED LICHEN

GROWTH-CURVE SITES OF SOLOMINA AND CALKIN (2003), CASCADE LAKE, AND OTHER LONG-TERM WEATHER STATIONS IN SOUTHWEST ALASKA

(DILLINGHAM AND BETHEL) Elevation

(m)

Mean annual temp.

(°C)

June, July, August (JJA) average

temp. (°C)

Mean annual precipitation

(mm)

Lichen growth-curve sites Brooks Range 350 -5.5 12.5 267 Seward Peninsula 4 -3.2 9.4 342 Wrangell-St. Elias 522 -5.7 13.3 201 Alaska Range 479 -2.9 12.6 239 Kenai (Homer) 27 3.0 11.1 518 Other climate stations

Bethel 36 1.4 12.0 430 Dillingham 20 1.0 11.7 647 Cascade Lake* 334 -4.0 8.7 Note: Climate data from the Spatial Climate Analysis Service (SCAS, 2005). *Based on daily averages at Cascade Lake from 2004-2006. Considered most similar to Seward Peninsula.

Page 79: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

71

TABLE 4. SUMMARY OF LICHEN SIZES ON MORAINES AND OTHER DEPOSITS Glacier Latitude Longitude Deposit type Five largest

(mm) All measured

(mm) n Single

largest

Site #

(Fig. 8) (°N) (°W) Mean ±S.D. Mean ±S.D. (mm)

Lichen year (based on Seward Pen. growth-curve with ~30 yr

colonization time) Cascade 1 60.1018 159.3493 left lateral* 39 2 33 4 23 42 71

Cascade 2 60.1009 159.3485 left lateral* 29 1 25 4 20 30 115

Cascade 3 60.1020 159.3446 right lateral 32 4 26 6 13 39 129

Cascade 4 60.1019 159.3460 right lateral 34 4 29 4 18 40 140

E. Inflow 5 60.1314 159.3603 moraine crest† 36 3 33 3 14 39 152

E. Inflow 6 60.1339 159.3653 debris flow† 26 1 23 4 8 28 10

E. Inflow 7 60.1343 159.3734 debris flow† 27 4 23 4 19 32 106

E. Inflow 8 60.1344 159.3647 debris flow† 28 4 25 4 12 32 111

Hungry Joe 9 60.1066 159.3235 stable outwash 23 4 23 4 3 33 91

Divided 10 60.1189 159.3215 slump 26 3 23 4 6 29 102

Chikuminuk 11 60.1495 159.3127 river terrace 33 9 27 9 10 49 135

Cascade 12 60.0942 159.3414 stable outwash lichen/no

lichen contact

N.A. N.A. 6 1 4 6 49

Cascade 13 60.0956 159.3426 stable outwash 16 6 13 5 10 27 70

*Shown in Figure 10A. †Shown in Figure 10B.

Page 80: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

72

TABLE 5. EQUILIBRIUM-LINE ALTITUDE (ELA) FOR GLACIERS IN THE CASCADE LAKE AREA Glacier Latitude Longitude 1957

length 1957 area

Head altitude

Toe altitude

Aspect 1957 ELA

Neoglacial ELA

ELA lowering

(see Fig. 8) (°N) (°W) (km) (km2) (m) (m) (m) (m) (m)

Cascade Glacier 60.0917 159.3380 3.8 3.76 1250 550 N 980 945 35

Divided Cirque 60.0951 159.3119 1.7 1.24 1190 730 NW 900 890 10

Hungry Joe Cirque 60.1066 159.3196 0.4 3.17 1190 980 SW 1045 1030 15

East Inflow Cirque 60.1311 159.3574 0.5 3.08 1070 820 NW 1085 1060 25

Chikuminuk Glacier 60.1300 159.2973 3.5 5.98 1280 610 NW 965 945 20

Cirque 1 60.0816 159.2999 0.6 1.69 1190 1010 NE 1045 1040 5

Cirque 2 60.0826 159.2843 0.5 2.78 1100 940 SE 1010 965 45

Cirque 3 60.1398 159.3378 1.0 4.08 1160 670 S 820 795 25

Average 980 960 23

Std. deviation 90 80 13

Page 81: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

73

TABLE 6. LOCATION AND OTHER INFORMATION FOR LAKE CORES RECOVERED FROM CASCADE LAKE

Core Latitude (°N)

Longitude (°W)

Water depth (m)

Length (m)

Notes

04CA-01 60.1432 159.4167 24.2 4.68 04CA-01A 60.1432 159.4167 24 0.20 Sub-sampled in field

04CA-01B 60.1432 159.4167 24 0.46

05CA-01A 60.1435 159.4166 23 0.20 Sub-sampled in field

05CA-01B 60.1435 159.4166 23 0.46

05CA-01C 60.1435 159.4166 23 0.35

04CA-02 60.1327 159.3988 34.5 4.37 Some material lost from base during core recovery

04CA-02A 60.1327 159.3988 34 0.20 sub-sampled in field

04CA-02B 60.1327 159.3988 34 0.38

05CA-02D 60.1333 159.3995 34 0.20 Sub-sampled in field

05CA-02E 60.1333 159.3995 34 0.38

05CA-02F 60.1333 159.3995 34 0.31

04CA-03 60.1371 159.4039 52.4 6.03

04CA-03B 60.1371 159.4039 52 0.30

Note: Surface cores indicated by letter designations.

Page 82: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

74

TABLE 7. TEPHRA DESCRIPTIONS FROM CORE 02 AND THE LAKESHORE

TEST PIT Basal depth (cm)

Thickness

(cm)

Glass type and color Bedding Contacts Notes

Core 02

2.59 1.1 Pink to cream bubble-walled shards

N.A. Sharp upper and lower

Not located in Core 01

2.85 15.0 Fine gray bubble-walled shards, coarse white pumice, significant microlitic glass shards

Reversely graded at base

Gradational upper and lower

Inferred to be the Aniakchak tephra (3500 cal yr BP)

3.24 3.0 Rounded with bubble wall junctions, clear to brown

N.A. Sharp upper and lower

Diatoms present

3.33 2.5 Brown glass shards N.A. Gradational upper, sharp lower

3.41 2.5 Brown glass shards N.A. Gradational upper, sharp lower

Test pit

4.30 2.0 Gray to white bubble wall shards

Preserved in pods

N.A. Overlies buried soil with a 14C age of 505 ± 20 cal yr BP

15.40 5.4 Orange/beige bubble wall shards with white pumice

Reversely graded at base

Gradational top, sharp bottom

Note: test pit illustrated in Figure 14.

Page 83: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

75

TABLE 8. CORRELATIONS BETWEEN BSi FLUX AND METEOROLOGICAL AND PACIFIC OCEAN CLIMATE INDICES Correlated parameter* Unsmoothed 5-year running avg. Source† n-years R p-value n-years R p-values Annual precipitation 19 0.35 0.144 18 0.44 0.437 1

May precipitation 19 0.03 0.892 14 -0.10 0.733 1

Mean annual temp 32 0.14 0.447 26 0.11 0.576 1

Mean DJF temp 26 -0.04 0.828 21 -0.35 0.120 1

Mean JJA temp 21 0.26 0.260 17 0.36 0.159 1

Mean may temp 31 0.54 0.002 25 0.70 9.16 x 10-5 1

Nuyakuk River discharge 26 0.73 2.22 x 10-5 23 0.71 1.27 x 10-4 2

PDOI 51 0.31 0.028 49 0.42 0.003 3

NPI 52 0.04 0.760 50 0.17 0.229 4

DJFM NPI 52 -0.42 0.002 50 -0.42 0.002 4

ALPI 51 0.40 0.004 50 0.68 1.37 x 10-7 5

PNI 53 0.32 0.022 52 0.41 0.003 6 *DJF = December, January, February; JJA = June, July, August; PDOI = Pacific Decadal Oscillation Index; NPI = North Pacific Index; DJFM = December, January, February, March; ALPI = Aleutian Low Pacific Index; PNI = Pacific Northwest Index. †(1) WRCC, 2006 and NCDC, 2005; (2) NWIS, 2006; (3) Mantua et al., 1997; (4) NCAR, 2006; (5) Beamish et al., 1997; (6) Ebbesmeyer et al., 1995.

Page 84: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

76

Figure 1: Periods of Holocene climate change. Columns denote the six periods of rapid climate change (polar cooling, tropical aridity)

proposed by Mayewski et al. (2004). Black boxes are centered over periods of climate cooling determined by various authors (left

column), and white boxes indicate mild, warm periods recognized by O’Brien et al. (1995).

Page 85: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

77

Figure 2: Location of the Ahklun Mountains and features discussed in this thesis

(medium shading indicates elevations above 200 m). Lakes are indicated by numbers: (1)

= Cascade Lake; (2) = Waskey Lake; (3) = Arolik Lake; (4) = Little Swift Lake; (5) =

Farewell Lake; (6) = Iceberg Lake. Other locations indicated by letter designations: (AR)

= Alaska Range; (B) = Bethel; (BR) = Brooks Range; (BB) = Bristol Bay; (D) =

Dillingham; (KM) = Kenai Mountains; (KB) = Kuskokwim Bay; (NR) = Nuyakuk River;

(SP) = Seward Peninsula; (WSEM) = Wrangell-St. Elias Mountains.

Page 86: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

78

Figure 3: Cascade Lake area. 1983 aerial photograph with arrows denoting inflows to

Cascade Lake; the western inflow is not glacier-fed. M1=Mooring 1, M2=Mooring 2,

AT=Air temperature station, GT=Glacier temperature station.

Page 87: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

79

Figure 4: Instrumentation deployed at Cascade Lake. (A) Sediment trap and water-

temperature logger mounted on mooring line. (B) Installing air-temperature logger

housed in a solar-radiation shield in a stand of willow along the shoreline of Cascade

Lake.

Page 88: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

80

Figure 5: Lithologic descriptions of the sediments collected in traps from Cascade Lake on August 10, 2005 and July 19, 2006.

Lithostratigraphy of bottom sediment traps from both mooring sites with physical and biological data (listed in Table 2).

Page 89: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

81

Figure 6: Daily average water temperature in Cascade Lake 2004-2006. Bottom loggers

were positioned ~1 m above the lake bottom; top loggers were ~3 m below the water

surface.

Page 90: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

82

Figure 7: Comparisons between average daily temperature at Cascade Lake, Dillingham, and Bethel, AK. (A) Differences between

sites. (B) Relationship between sites; note the slight seasonality effect shown by the increased offset from the 1:1 line for the coldest

winter temperatures.

Page 91: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

83

Figure 8: Cascade Lake area showing glaciers referenced by informal names

(CA=Cascade Glacier, C1=Cirque 1, C2=Cirque 2, DIV=Divided Glacier, HJ=Hungry

Joe, CH=Chikuminuk Glacier, EI=East Inflow Cirque, C3=Cirque 3). Based on 1:63,360

Bethel 1979 topographic map with corrections for ice extent based on 1957 aerial

photographs.

Page 92: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

84

Figure 9: Hypsometric curves used in the calculation of equilibrium-line altitudes (ELA)

for eight glaciers in the study area. An accumulation-area ratio (AAR) of 0.6 (60% of the

glacier in the accumulation zone, shown by the line at 40% of the glacier area) was used

for ELA calculations. ELA calculations (and especially ∆ELA) are insensitive to AAR

changes as shown by the relatively flat slopes between AAR values of 0.5-0.7 (shown by

the box).

Page 93: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

85

Figure 10: Little Ice Age (LIA) moraines in the Cascade Lake drainage. Dotted lines highlight moraine crests and trim lines. (A) Left

and right lateral LIA moraines of Cascade Glacier. Deposits in the foreground are higher than those in the distance. (B) The East

Inflow cirque valley. The ridge crest closest to the cirque (dotted line) is the LIA moraine; the fan-shaped feature marked by ridges

below are likely debris-flow deposits. Numbers in white denote lichen stations. Photos taken in August 2005.

Page 94: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

86

Figure 11: Growth curves used to derive the ages of measured lichens in the Cascade

Lake region (from Solomina and Calkin, 2003). The arrows show the average of the five

larges lichen diameters on moraines in the Cascade Lake area. Based on Seward

Peninsula curve, their ages average ~1860 AD assuming a ~30-yr colonization time.

Page 95: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

87

Figure 12: Retreat rates for Cascade Glacier since the Little Ice Age based on aerial

photography (1984 and 1957) and lichen stations (1935 and 1860; see text).

Page 96: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

88

Figure 13: Bathymetry of Cascade Lake with locations of coring sites. The main inflow

is in the south and the outflow is to the north.

Page 97: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

89

Figure 14: Top 15 cm of Core 02 showing the relationship between the core surface

stratigraphy, X-radiograph stratigraphy, and the measurement of lamination thicknesses.

Lines connect points identified as being the same on the core face and X-radiograph

stratigraphy.

Page 98: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

90

Figure 15: Acoustic profile image in the vicinity of core site 02. The white box denotes

the approximate location and length of Core 02. The acoustic signal is degraded at the

approximate base of Core 02, but stratigraphy deeper than this was observed in other

profiles and below Core 01.

Page 99: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

91

Figure 16: Lithologic logs of three cores recovered from Cascade Lake.

Page 100: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

92

Figure 17: Correlations between surface and long cores based on marker beds (dashed lines) in (A) Core 01 and (B) Core 02. Depths

for surface core refer to the sediment-water interface (lake bottom), and are within 0.5 cm of correlative depths in long cores.

Page 101: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

93

Figure 18: Test-pit wall on the northern shore of Cascade Lake (Figure 13). * denotes

location of 14C sample dated at 505 cal yr BP.

Page 102: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

94

Figure 19: 210Pb and 137Cs profiles from Core 02A and Core 02. The 3rd-order

polynomial was used to assign ages to the upper 12 cm of Core 02. Gray box shows the

intersection of 137Cs peak with ages derived from the 210Pb profile.

Page 103: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

95

Figure 20: Geochemistry data from the 2.85 m ash in Core 02 and published data sets of

known Aniakchak tephras. Data listed in appendix A-2.

Page 104: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

96

Figure 21: Scanning electron microscope images of tephra shards from (A) the 2.85 m tephra in Core 02 and (B) a known Aniakchak

tephra from Arolik Lake.

Page 105: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

97

Figure 22: Age model for Core 02. The two models are based on the maximum depth

range of Tephra B candidates in Core 02 (see text for explanation).

Page 106: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

98

Figure 23: Magnetic susceptibility (MS) profiles for Cores 01, 02 and 03 from Cascade

Lake. * indicates MS values that exceed 90 x 10-6 SI and dashed lines show correlations

between tephras.

Page 107: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

99

Figure 24: Organic-matter (OM) content and bulk density profiles for the upper 2 m in

Core 01 and the entire Core 02.

Page 108: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

100

Figure 25: Lamination thickness in the top 1.4 m of Core 02. Of the 316 lamination

couplets identified, the average thickness is 0.42 cm (bold line).

Page 109: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

101

Figure 26: Biogenic silica profiles for Core 02. Inset shows detail of upper 14 cm, the

interval sampled at ~3 yr resolution and used to correlate with instrumental climate

records.

Page 110: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

102

Figure 27: Biogenic silica flux from Core 02 over the last 2200 cal yr BP with periods of

recognized warm climate from Iceberg Lake (Loso et al., 2006), Farewell Lake (Hu et al.,

2001), and Jellybean Lake (Anderson et al., 2005).

Page 111: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

103

Figure 28: Core 02 proxy data compared to climate events of the Yukon (Jellybean Lake; Anderson et al., 2005) and Waskey Lake

(Levy et al., 2004). OM = organic matter; BSi = biogenic silica. Gray bars indicate the Holocene Thermal Maximum (7000 cal yr

BP), Neoglaciation onset in the Ahklun Mountains (3000 cal yr BP) and Medieval Warm Period (~1000 cal yr BP).

Page 112: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

104

TABLE A-1. 210Pb AND 137Cs DATA

Depth Unsupported 210Pb activity

137Cs activity Age of base

(cm) (dpm/g) (± s.d.) (dpm/g) (±s.d.) (years before 2004) Core 02A 0.5 9.16 0.53 5.48 1.5 3.80 0.37 10.95 2.5 4.29 0.26 17.58 3.5 3.88 0.23 27.31 4.5 1.88 0.17 35.90 5.5 1.31 0.19 43.22 6.5 0.45 0.16 47.38 7.5 0.46 0.16 51.13 8.5 0.51 0.15 55.88 9.5 1.30 0.24 70.45 10.5 0.57 0.17 88.45 11.5 0.48 0.14 113.58 12.5 0.22 0.14

N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. N.D. 148.35

Core 02 1.5 7.10 0.78 0.30 0.13 2.86 2.0 8.28 0.57 0.56 0.08 6.74 2.5 6.27 0.69 0.71 0.10 10.04 3.0 6.52 0.54 0.60 0.08 13.80 3.5 5.70 0.81 0.83 0.13 17.53 4.0 3.37 0.78 0.73 0.13 20.01 4.5 3.29 0.57 1.22 0.10 22.51 5.0 3.51 0.58 1.14 0.10 25.59 5.5 2.28 0.78 1.78 0.14 27.71 6.0 2.33 0.59 1.95 0.10 30.21 6.5 2.92 0.73 5.58 0.16 33.43 7.0 3.27 0.70 4.20 0.14 42.11 8.5 1.90 0.59 1.10 0.10 52.51 10.0 2.53 0.53 0.30 0.10 87.27 13.0 1.14 0.51 0.00 0.1 153.57

Page 113: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

105

TABLE A-2. MAJOR ELEMENT GEOCHEMISTRY OF THE 2.85 m TEPHRA IN CORE 02

Point Na2O MgO Al2O3 SiO2 P2O5 K2O CaO TiO2 MnO FeO Total

1 2.397 0.489 15.370 72.958 0.069 2.651 1.706 0.439 0.152 2.392 98.623 2 2.228 0.488 15.429 72.122 0.057 2.415 1.767 0.439 0.094 2.385 97.424 3 3.302 2.605 16.468 59.778 0.784 1.409 5.906 1.306 0.186 6.800 98.545 4 1.290 0.489 15.514 72.879 0.096 2.225 1.797 0.394 0.097 2.357 97.137 5 2.708 3.220 16.466 57.271 0.807 1.131 7.146 1.443 0.194 8.197 98.583 6 1.525 1.323 16.211 64.919 0.538 1.449 3.839 1.093 0.090 3.760 94.748 7 0.997 0.464 15.582 74.064 0.092 1.718 1.820 0.445 0.105 2.434 97.722 8 1.645 2.225 16.145 62.412 0.577 1.508 5.334 1.253 0.142 6.174 97.420 9 1.529 2.784 16.519 58.862 0.754 1.191 6.361 1.409 0.178 7.299 96.888 10 1.089 0.439 15.242 72.922 0.103 1.523 1.767 0465 0.085 2.217 95.853 11 1.665 2.406 16.665 61.372 0.747 1.311 5.706 1.328 0.159 6.429 97.787 12 2.304 2.555 16.635 60.803 0.655 1.229 6.249 1.308 0.223 7.166 99.126 13 1.988 2.920 16.026 56.595 0.658 1.114 6.398 1333 0.213 8.197 95.443 14 1.007 0.479 15.537 73.401 0.137 1.533 1.854 0.435 0.121 2.417 96.923 15 2.141 2.793 16.298 58.920 0.644 1.119 6.287 1.323 0.159 7.402 97.085 16 0.737 0.363 12.259 56.077 0.057 1.075 1.545 0.379 0.090 2.120 74.702 17 0.743 0.406 12.763 61.909 0.050 0.928 1.591 0.395 0.039 2.240 81.064 18 1.851 1.711 16.973 64.354 0.435 1.444 4.587 1.091 0.156 4.729 97.331 19 2.188 0.425 15.680 74.205 0.028 1.957 1.692 0.484 0.110 2.428 99.196 20 1.334 0.428 15.369 73.521 0.147 1.788 1.713 0.470 0.141 2.304 97.214 21 3.169 3.365 16.294 57.525 0.603 0.875 7.251 1.281 0.152 8.807 99.322 22 2.668 2.849 16.595 59.517 0.726 1.024 6.049 1.358 0.199 7.033 98.017 23 3.193 2.899 16.655 58.888 0.758 0.890 6.702 1.433 0.199 7.032 98.650 24 1.587 0.839 15.478 71.489 0.034 1.629 2.695 0.570 0.077 3.061 97.459 25 1.256 0.491 15.828 73.831 0.034 1.636 1.990 0.445 0.114 2.366 97.991 26 2.564 2.577 16.963 60.082 0.625 1.012 5.928 1.528 0.117 6.587 98.042 27 1.349 0.413 15.524 73.514 0.000 1.683 1.880 0.460 0.124 2.556 97.504 28 1.426 0.463 14.704 70.288 0.050 1.432 1.763 0.434 0.119 2.213 92.892 29 1.086 0.446 14.908 71.677 0.069 1.160 1.851 0.439 0.094 2.287 94.017 30 2.796 3.113 16.287 57.436 .736 0.790 6.657 1.311 0.196 7.468 96.789 31 1.277 0.486 15.331 73.576 0080 1.390 1.840 0.429 0.085 2.404 96.898 std. 2.587 5.051 12.431 50.532 0.509 0.729 9.472 3.900 0.198 13.473 98.881

Page 114: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

106

TABLE A-3. MAGNETIC SUSCEPTIBILITY DATA FOR CORE 02

0 20 40 60 80 100 120 140 160 180 200 0.0 27.6 33.8 27.0 27.0 23.7 27.0 24.6 28.1 27.0 23.5 0.5 27.9 27.8 30.9 26.0 28.0 24.9 26.0 22.8 26.4 28.0 25.3 1.0 26.9 27.2 28.4 24.9 27.8 26.0 28.0 22.2 26.7 26.0 31.2 1.5 26.0 33.4 27.5 31.7 27.7 24.0 26.0 20.2 32.0 26.1 26.9 2.0 23.0 31.5 30.6 26.5 26.5 26.0 25.0 15.3 29.0 24.3 21.7 2.5 23.0 29.7 29.7 24.3 29.4 31.0 26.0 16.4 28.0 23.5 22.5 3.0 27.0 33.9 28.9 26.2 28.2 29.0 25.0 23.5 28.0 24.6 24.4 3.5 19.0 34.2 30.0 28.0 28.7 29.0 24.2 23.6 27.0 26.8 26.2 4.0 22.0 34.3 29.0 23.0 26.4 28.0 24.4 23.6 25.2 26.0 21.8 4.5 26.0 30.5 29.0 25.0 25.0 28.0 24.6 22.7 17.6 26.0 20.6 5.0 35.0 26.7 36.0 28.0 25.7 25.0 23.7 23.8 25.5 26.0 21.5 5.5 29.0 27.9 27.0 30.0 27.3 24.0 23.9 23.9 28.7 24.0 24.3 6.0 28.8 24.0 25.0 28.0 29.0 28.2 26.7 20.3 26.4 25.0 24.2 6.5 32.0 25.0 26.0 25.0 28.0 24.3 26.6 20.5 24.5 24.6 28.7 7.0 31.3 25.0 30.0 24.0 28.0 24.5 27.4 23.6 27.6 24.5 25.6 7.5 31.5 31.0 26.0 23.0 27.0 23.7 27.3 21.7 31.7 23.4 24.5 8.0 33.8 32.0 23.5 23.0 28.0 25.8 28.2 22.9 29.8 21.3 26.4 8.5 29.1 33.0 26.6 23.2 28.0 25.1 30.0 28.1 26.9 22.1 31.2 9.0 27.3 30.0 33.7 24.3 26.0 29.3 29.0 25.1 30.0 25.0 29.7 9.5 29.5 24.0 25.7 25.5 25.0 31.5 27.0 18.2 29.0 25.0 30.6 10.0 27.6 24.0 26.8 25.7 23.0 25.7 28.0 18.5 29.0 25.0 27.4 10.5 27.8 28.0 27.9 27.8 26.0 26.8 26.0 18.9 29.0 26.0 23.3 11.0 28.1 28.2 28.9 26.0 26.0 26.1 27.0 19.8 31.0 23.0 25.1 11.5 26.3 29.3 26.7 26.0 24.0 28.3 26.0 16.0 32.0 24.9 27.6 12.0 24.5 29.5 28.5 23.0 25.0 25.5 25.0 24.0 31.0 24.7 27.5 12.5 25.7 29.7 29.4 23.0 23.0 25.6 27.0 27.0 27.0 26.5 25.4 13.0 20.9 31.9 27.2 22.0 26.0 26.8 24.0 26.0 26.0 23.4 25.2 13.5 20.2 32.0 27.8 27.0 27.0 24.0 23.0 27.0 27.0 23.2 26.1 14.0 22.4 29.0 29.6 23.0 26.0 24.0 23.0 24.3 30.8 21.8 26.8 14.5 24.6 29.0 31.5 24.0 27.0 24.0 26.0 25.6 27.6 21.6 26.7 15.0 28.7 30.0 29.3 26.0 27.0 25.0 25.0 27.0 26.5 22.5 31.5 15.5 28.9 27.0 27.1 25.0 30.0 25.0 24.0 27.3 26.3 23.3 33.3 16.0 29.8 24.1 25.8 30.1 29.2 26.0 24.4 25.7 29.1 24.1 31.2 16.5 26.7 28.3 29.6 25.3 26.3 23.0 24.5 24.5 27.0 24.5 27.0 17.0 25.5 30.5 27.5 24.5 27.5 25.0 25.6 25.0 29.0 23.4 26.0 17.5 25.4 29.6 27.3 25.6 27.6 27.0 21.8 28.5 27.0 24.3 25.0 18.0 29.2 29.8 24.2 26.8 26.8 20.0 22.9 30.0 28.0 24.2 27.0 18.5 27.1 31.4 25.0 25.0 24.3 29.0 22.2 26.4 26.0 25.1 27.0 19.0 29.3 32.5 28.0 28.0 23.4 28.0 21.3 26.3 27.0 27.8 27.0 19.5 27.4 34.6 27.0 29.0 25.6 28.0 24.5 26.6 28.0 24.7 29.0

Page 115: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

107

220 240 260 280 300 320 340 360 380 400 420 0.0 29.0 30.4 55.5 289.5 22.0 21.2 17.0 15.6 17.0 27.2 23.0 0.5 25.0 28.3 58.4 227.4 21.0 18.4 16.0 16.5 18.0 23.3 23.0 1.0 26.0 27.1 63.2 204.2 20.0 15.5 17.0 15.4 17.0 15.5 31.0 1.5 28.9 27.8 65.7 241.9 19.0 14.7 17.0 15.3 17.0 15.7 30.0 2.0 29.7 25.7 73.6 220.8 17.0 14.9 16.0 14.1 15.0 21.8 26.0 2.5 25.5 25.5 100.4 274.7 16.0 16.2 17.0 16.0 16.1 26.2 26.7 3.0 22.3 27.4 136.3 299.6 16.0 19.4 16.0 17.0 17.3 25.4 25.8 3.5 21.2 31.2 115.1 332.5 17.0 21.5 16.0 16.0 17.4 23.5 26.8 4.0 22.2 29.8 119.8 312.5 18.0 21.7 16.0 15.0 18.6 23.7 30.9 4.5 24.4 25.7 106.7 314.3 17.0 20.9 17.0 14.0 19.8 19.8 30.9 5.0 25.5 23.5 141.5 287.2 16.0 17.3 18.0 14.1 21.9 19.1 30.2 5.5 24.7 23.4 156.3 143.2 16.0 16.6 17.0 14.3 21.7 22.3 32.3 6.0 23.8 22.2 161.2 120.1 17.0 17.0 16.0 16.5 20.5 16.5 34.5 6.5 23.0 25.7 128.1 131.0 14.0 18.3 15.0 19.6 21.3 21.6 33.7 7.0 22.0 31.6 117.3 148.0 13.0 18.7 17.0 21.8 20.2 21.8 24.9 7.5 22.0 32.4 143.4 87.0 11.7 20.0 19.0 15.0 21.0 19.8 21.0 8.0 24.0 34.3 161.6 43.0 16.9 23.0 20.0 15.0 21.0 16.7 19.0 8.5 28.0 29.1 131.8 25.0 17.2 25.0 15.0 15.0 21.0 22.5 17.0 9.0 32.2 31.0 147.0 22.8 17.4 27.0 15.0 18.0 18.0 19.4 17.0 9.5 26.4 40.0 121.0 21.7 16.7 34.0 12.0 23.0 20.0 17.2 28.0 10.0 27.5 40.0 52.0 21.5 18.2 39.2 13.0 20.0 20.4 23.0 28.0 10.5 27.7 33.0 86.0 23.3 19.3 67.4 14.0 16.0 19.5 17.0 31.0 11.0 30.8 28.0 113.0 23.2 16.5 104.5 13.0 15.0 20.6 19.0 33.0 11.5 25.2 27.7 105.7 22.5 17.6 27.7 14.0 15.0 21.7 30.0 30.0 12.0 23.4 27.6 127.6 22.4 18.9 17.9 12.0 19.0 19.9 30.0 29.0 12.5 24.5 29.5 144.5 22.6 18.3 16.0 13.0 15.9 21.2 26.8 23.0 13.0 21.7 36.3 109.3 21.2 18.6 16.0 13.0 16.7 19.4 17.6 24.0 13.5 23.9 40.2 79.2 20.1 19.0 17.0 13.0 18.5 19.5 20.5 26.0 14.0 28.0 77.6 95.0 19.8 20.3 18.0 13.0 16.3 20.7 24.3 24.0 14.5 27.0 114.3 148.8 19.7 24.6 17.0 13.0 15.2 19.9 22.1 23.0 15.0 27.0 71.3 159.6 20.5 25.2 17.0 15.0 16.1 21.0 25.8 23.8 15.5 27.0 48.2 164.4 24.4 24.3 17.0 15.0 20.3 20.0 20.7 22.6 16.0 27.0 37.1 125.1 20.2 22.5 19.0 15.0 21.4 22.0 18.5 28.4 16.5 28.2 46.0 150.0 21.0 23.6 23.0 16.0 16.6 21.0 22.4 26.2 17.0 29.4 51.0 129.0 22.0 25.8 24.0 14.0 17.8 20.0 25.2 17.5 27.6 49.0 130.0 23.0 24.2 21.9 14.0 17.0 20.1 22.0 18.0 24.7 47.0 185.0 23.0 22.3 22.7 13.0 24.0 20.3 26.0 18.5 27.9 39.0 202.0 22.0 21.5 25.5 13.0 26.0 18.5 29.0 19.0 28.6 48.8 270.8 21.0 24.7 22.3 11.0 22.0 24.7 21.0 19.5 27.5 53.7 342.7 22.0 25.9 20.1 14.0 16.0 20.8 23.0 Note: The sums of the column and row headings indicate the depth for the magnetic susceptibility readings. All magnetic susceptibility readings reported in 10-6 SI units.

Page 116: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

108

TABLE A-4. PERCENT ORGANIC MATTER DATA FOR CORE 02

0 40 80 120 160 200 240 280 0 2.97 3.34 3.37 3.33 3.48 3.40 N.D. 1 2.36 3.27 3.01 3.53 3.25 4.02 3.74 N.D. 2 2.43 2.96 2.99 3.31 3.20 4.00 3.90 2.39 3 2.25 3.18 2.91 3.33 3.33 3.59 3.69 N.D. 4 2.25 3.25 2.93 3.65 3.43 3.45 3.79 N.D. 5 2.29 3.17 2.95 3.65 3.40 3.52 3.79 N.D. 6 2.40 3.24 2.74 3.39 3.35 3.69 3.92 N.D. 7 2.15 3.17 2.64 3.29 3.66 3.44 3.80 2.66 8 2.17 3.30 2.94 3.19 3.52 3.48 3.98 N.D. 9 2.17 3.15 2.95 3.20 3.41 3.80 3.92 3.63 10 2.49 3.40 3.31 3.40 3.17 3.65 4.18 3.93 11 2.74 3.08 3.07 3.88 3.28 4.08 4.45 3.43 12 2.84 2.84 3.04 3.45 3.03 3.87 3.76 3.59 13 2.95 2.99 2.96 3.50 3.46 3.62 3.71 3.56 14 3.96 2.95 2.93 3.35 3.43 3.43 3.50 3.64 15 3.36 3.03 2.95 3.39 3.19 3.71 3.01 3.53 16 2.62 3.25 2.95 3.33 3.41 3.79 3.42 3.41 17 2.89 3.15 2.89 3.51 3.16 3.62 3.96 3.50 18 2.78 3.36 2.91 3.27 3.17 3.51 3.92 3.49 19 2.69 3.26 3.18 3.44 3.48 3.43 4.54 3.55 20 2.93 3.49 3.15 3.36 3.39 2.96 4.08 3.59 21 2.56 3.00 2.81 3.38 3.57 3.88 3.82 3.54 22 2.82 3.23 2.97 3.46 3.64 3.61 3.70 3.80 23 2.66 3.41 2.96 3.30 3.32 3.58 3.50 3.44 24 3.00 3.93 3.11 3.45 3.05 3.61 3.53 3.89 25 3.45 2.88 3.12 3.68 3.08 3.47 2.86 N.D. 26 3.43 3.20 3.31 3.25 3.22 3.52 N.D. N.D. 27 3.04 3.19 3.30 3.31 3.51 3.06 N.D. N.D. 28 2.49 3.32 3.49 3.26 3.41 3.24 N.D. N.D. 29 2.92 3.07 3.33 3.44 3.34 3.30 N.D. N.D. 30 2.64 3.08 3.44 3.32 3.63 3.60 N.D. N.D. 31 2.61 3.14 3.21 N.D. 3.55 3.68 N.D. N.D. 32 2.73 3.35 3.53 3.36 3.44 3.43 N.D. N.D. 33 2.41 3.31 3.49 3.27 3.73 3.36 N.D. N.D. 34 3.22 3.32 3.77 3.32 3.46 3.70 N.D. N.D. 35 2.84 3.30 3.71 3.29 3.62 3.90 N.D. N.D. 36 3.26 3.00 3.71 3.46 3.38 3.86 N.D. N.D. 37 2.83 3.04 3.63 3.59 3.49 4.12 N.D. N.D. 38 3.07 3.11 3.34 3.84 3.82 3.65 N.D. N.D. 39 2.72 3.30 3.61 3.01 3.70 3.51 N.D. N.D.

Page 117: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

109

300.0 340.5 380.5 420.5

0.0 N.D. 4.61 3.88 1.86 1.5 N.D. 4.51 3.81 2.07 2.5 N.D. 4.59 3.80 1.87 3.5 N.D. 4.63 3.89 1.55 4.5 N.D. 4.51 3.72 1.87 5.5 3.93 4.41 3.10 1.74 6.5 3.66 4.62 3.48 1.44 7.5 3.78 4.28 3.27 1.43 8.5 3.89 4.31 3.39 2.38 9.5 3.82 4.24 3.12 2.26 10.5 3.64 3.19 3.36 1.74 11.5 3.74 4.39 3.01 12.5 3.91 4.38 2.94 13.5 3.95 4.38 3.00 14.5 3.91 4.43 2.95 15.5 3.63 4.40 2.99 16.5 3.75 4.40 3.00 17.5 3.85 4.37 2.61 18.5 3.74 4.38 2.64 19.5 3.93 3.67 2.75 20.5 3.93 4.12 2.59 21.5 4.07 4.30 3.13 22.5 4.16 4.23 2.80 23.5 4.05 4.04 2.57 24.5 4.06 4.05 2.69 25.5 4.23 4.03 2.58 26.5 4.39 4.27 2.44 27.5 4.44 3.89 2.54 28.5 4.29 4.00 2.77 29.5 4.35 4.04 2.81 30.5 4.42 3.95 2.54 31.5 4.44 3.88 2.31 32.5 4.69 3.75 1.97 33.5 4.28 4.01 2.73 34.5 4.39 3.82 2.46 35.5 4.54 4.07 2.60 36.5 4.74 3.83 2.42 37.5 4.63 4.11 2.40 38.5 4.72 4.07 2.41 39.5 4.64 3.60 2.12 Note: The sums of the column and row headings indicate the depth for the loss-on-ignition measurements.

Page 118: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

110

TABLE A-5. BIOGENIC SILICA DATA FOR CORE 02

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 0.60 1.75 0.23 0.80 2.00 0.82 1.00 1.92 1.35 1.20 2.47 2.05 1.40 2.06 1.99 1.60 2.68 2.94 1.80 2.67 3.30 2.00 2.92 4.01 2.10 3.23 4.39 2.40 2.44 3.14 2.60 2.47 3.04 2.80 2.66 3.12 3.00 2.35 2.62 3.20 2.32 2.59 3.40 2.10 2.32 3.60 2.37 2.61 3.80 1.99 2.17 4.00 2.72 2.93 4.20 2.37 2.61 4.40 2.30 2.58 4.60 2.14 2.45 4.80 1.90 2.20 5.00 2.23 2.61 5.20 2.08 2.42 5.40 1.99 2.29 5.60 2.04 2.32 5.80 2.22 2.49 6.00 2.03 2.24 6.20 1.62 1.63 6.40 2.35 2.21 6.60 2.36 2.06 6.80 2.01 1.62 7.00 2.42 1.80 7.20 1.64 1.23 7.40 2.37 1.74 7.60 1.78 1.28 7.80 2.23 1.56 8.00 2.07 1.41 8.20 1.97 1.34 8.40 1.81 1.24 8.60 2.01 1.37 8.80 1.90 1.29 9.00 1.86 1.26

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 9.20 1.75 1.16 9.40 2.01 1.30 9.60 2.06 1.31 9.80 2.00 1.25 10.00 2.20 1.34 10.20 2.09 1.18 10.40 1.99 1.04 10.60 1.81 0.87 10.80 1.70 1.77 11.00 1.64 1.63 11.30 1.76 1.69 11.60 2.08 1.92 11.80 2.05 1.84 12.00 2.03 1.78 12.40 1.96 1.76 12.60 2.62 2.38 12.80 1.86 1.71 13.00 2.03 1.89 13.20 2.20 2.18 13.40 2.09 2.19 13.60 1.93 2.14 13.80 2.00 2.34 14.00 1.88 2.32 14.20 1.96 2.36 14.40 1.96 2.29 15.75 1.88 1.79 16.25 2.07 1.97 16.75 2.12 2.26 17.25 1.85 1.97 17.75 2.14 2.11 18.25 2.11 2.09 18.75 2.25 2.58 19.25 1.63 1.87 19.75 2.20 2.64 20.25 2.09 2.50 20.75 1.93 2.11 21.25 1.78 1.94 21.75 1.81 1.94 22.25 1.88 2.01 22.75 2.06 2.08 23.25 1.47 1.49 23.75 2.17 2.24 24.25 1.84 1.90 24.75 2.47 2.39 25.25 2.19 2.11

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 25.75 1.70 1.87 26.25 1.57 1.73 26.75 1.70 1.27 27.25 1.49 1.12 27.75 1.89 1.75 28.25 2.09 1.95 28.75 2.32 2.23 29.25 2.20 2.12 29.75 1.91 1.49 30.25 2.04 1.59 30.75 1.81 1.64 31.25 2.07 1.88 31.75 2.02 2.12 32.25 1.85 1.95 32.75 1.96 1.37 33.25 1.79 1.25 33.75 2.08 1.92 34.25 1.97 1.82 34.75 1.89 1.88 35.25 1.58 1.58 35.75 2.00 1.74 36.25 1.98 1.72 36.75 1.73 1.78 37.25 2.99 3.07 37.75 1.91 1.98 38.25 1.54 1.60 38.75 1.82 2.11 39.25 1.92 2.23 39.75 1.92 1.91 40.25 1.90 1.89 40.75 1.73 1.48 41.25 1.82 1.56 41.75 1.80 1.63 42.25 1.69 1.53 42.75 1.88 1.54 43.25 1.73 1.42 43.75 1.86 2.43 44.25 1.81 2.36 44.75 1.89 1.47 45.25 1.89 1.46 45.75 1.86 1.59 46.25 1.95 1.67 46.75 1.63 1.66 47.25 2.25 2.29 47.75 1.93 2.05

Page 119: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

112

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 48.25 1.95 2.06 48.75 1.82 2.13 49.25 1.96 2.30 49.75 1.47 2.06 50.25 1.71 2.39 50.75 1.91 1.79 51.25 1.88 1.76 51.75 1.78 1.72 52.25 1.76 1.71 52.75 2.04 2.18 53.25 1.61 1.72 53.75 1.69 2.32 54.25 1.64 2.25 54.75 1.74 1.67 55.25 1.85 1.77 55.75 1.53 1.79 56.25 1.56 1.83 56.75 1.64 1.86 57.25 1.28 1.45 57.75 2.25 2.16 58.25 1.76 1.70 58.75 1.91 2.26 59.25 1.59 1.88 59.75 1.57 1.83 60.25 1.81 2.11 60.75 1.98 2.04 61.25 1.88 1.93 61.75 1.79 1.72 62.25 1.91 1.84 62.75 1.88 2.35 63.25 1.89 2.37 63.75 1.79 2.16 64.25 1.36 1.64 64.75 1.85 2.01 65.25 1.74 2.25 65.75 1.71 1.15 66.25 1.76 1.19 66.75 1.74 1.34 67.25 1.37 1.05 67.75 1.76 1.96 68.25 2.06 2.29 68.75 2.09 2.24 69.25 1.78 1.91 69.75 1.96 2.05

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 70.25 1.95 2.04 70.75 2.02 2.36 71.25 1.80 2.10 71.75 1.34 1.41 72.25 1.30 1.36 72.75 2.31 1.65 73.25 1.40 1.00 73.75 2.07 1.85 74.25 2.11 1.89 74.75 0.95 0.88 75.25 1.82 1.68 75.75 1.56 1.65 76.25 2.02 2.14 76.75 1.97 1.88 77.25 1.84 1.75 77.75 1.82 1.42 78.25 2.15 1.69 78.75 1.79 1.90 79.25 1.97 2.09 79.75 1.90 2.55 80.25 1.87 2.51 80.75 2.17 2.24 81.25 1.84 1.90 81.75 2.05 2.87 82.25 1.76 2.47 82.75 1.81 2.07 83.25 1.90 2.17 83.75 1.96 2.04 84.25 1.86 1.94 84.75 1.84 2.14 85.25 1.73 2.01 85.75 1.98 1.72 86.25 2.08 1.81 86.75 2.09 2.22 87.25 2.20 2.35 87.75 1.96 2.37 88.25 1.94 2.35 88.75 1.85 2.42 89.25 1.95 2.54 89.75 2.00 1.73 90.25 2.18 1.88 90.75 1.97 1.82 91.25 2.06 1.90 91.75 2.01 1.59 92.25 2.20 1.74 92.75 2.12 2.54

Depth BSi BSi Flux s(cm) (%) (mg·cm-2

·yr-1) 93.25 2.09 2.51 93.75 2.16 2.84 94.25 2.01 2.65 94.75 2.26 3.05 95.25 1.69 2.29 95.75 2.65 3.00 96.25 2.76 3.13 96.75 2.46 3.09 97.25 2.41 3.02 97.75 2.44 2.87 98.25 2.81 3.30 98.75 2.76 2.74 99.25 2.54 2.52 99.75 2.23 2.22 100.25 2.46 2.45 100.75 2.64 2.63 101.25 2.52 2.51 101.75 2.66 3.10 102.25 2.39 2.78 102.75 2.51 2.85 103.25 2.49 2.82 103.75 2.40 2.83 104.25 2.15 2.53 104.75 2.37 2.34 105.25 2.70 2.66 105.75 2.95 2.96 106.25 2.89 2.90 106.75 2.71 2.92 107.25 2.51 2.70 107.75 2.75 2.36 108.25 1.57 1.34 108.75 2.14 2.41 109.25 2.54 2.86 109.75 2.55 2.30 110.25 2.65 2.38 110.75 2.51 2.68 111.25 2.12 2.25 111.75 2.63 2.86 112.25 2.68 2.91 112.75 2.60 2.47 113.25 2.66 2.53 113.75 2.71 2.31 114.25 2.59 2.21 114.75 2.16 2.54 115.25 2.11 2.47 115.75 1.74 2.48

111

Page 120: Backup of Final - nau.edu · drainage basin of Cascade Lake, (3) identifying the major sedimentation patterns and fluctuations in Cascade Lake sediment cores, and (4) using proxy

112

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 116.25 1.90 1.79 117.25 1.56 1.76 117.75 1.60 1.58 118.25 1.82 1.79 118.75 1.64 1.80 119.25 1.84 2.03 119.75 1.88 2.01 120.25 1.81 1.94 120.75 1.54 1.64 121.25 1.40 1.50 121.75 1.50 1.44 122.25 1.46 1.40 122.75 1.55 1.49 123.25 1.98 1.90 123.75 1.65 1.33 124.25 1.52 1.23 124.75 1.72 1.58 125.25 2.60 2.39 125.75 1.90 1.83 126.25 1.83 1.77 126.75 1.81 1.41 127.25 1.65 1.29 127.75 1.79 2.00 128.25 1.80 2.00 128.75 1.75 1.59 129.25 1.69 1.54 129.75 1.68 1.79 130.25 1.75 1.87 130.75 1.73 1.46 131.25 1.67 1.41 131.75 1.77 1.65 132.25 1.85 1.72 132.75 1.90 1.74 133.25 1.83 1.67 133.75 1.75 1.10 134.25 1.70 1.07 134.75 1.67 1.55 135.25 1.68 1.56 135.75 1.03 0.95 136.25 1.09 0.97 136.75 1.04 0.82 137.25 1.41 1.10 137.75 1.22 1.42 138.75 1.29 1.53 139.25 1.17 1.38 139.75 1.51 1.29 140.25 1.17 1.00 140.75 1.25 1.41

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 141.25 1.45 1.63 141.75 1.42 1.21 142.25 1.32 1.12 142.75 1.13 1.03 143.25 1.24 1.12 143.75 1.08 0.82 144.25 1.19 0.90 144.75 1.21 1.26 145.25 1.33 1.38 145.75 1.33 1.80 146.25 1.29 1.75 146.75 1.15 0.95 147.25 1.26 1.04 147.75 1.27 1.29 148.25 1.23 1.25 148.75 1.42 1.26 149.25 1.32 1.17 149.75 1.47 1.19 150.25 1.45 1.16 150.75 1.35 1.09 156.25 1.29 1.03 161.25 1.67 1.21 166.25 1.33 1.17 171.25 1.76 1.32 176.25 1.84 1.42 181.25 1.72 1.41 186.00 1.92 1.39 191.25 1.76 1.53 196.25 1.76 1.19 201.25 1.53 1.28 206.25 1.91 1.52 211.25 2.19 1.45 216.25 2.30 1.72 221.25 1.59 1.12 226.25 2.20 1.22 231.25 2.18 2.17 236.25 3.06 2.28 241.25 4.09 2.30 246.25 4.07 1.99 251.25 4.86 3.60 256.25 3.39 1.82 260.25 3.80 2.71 260.75 5.26 3.83 261.25 5.41 3.94 261.75 4.33 4.03 262.25 3.87 3.60 262.75 4.68 2.86 263.25 4.57 2.79

Depth BSi BSi Flux (cm) (%) (mg·cm-2

·yr-1) 263.75 3.17 2.65 266.25 3.79 N.D. 288.25 2.75 0.42 289.25 2.89 0.44 290.25 1.85 0.25 291.25 1.71 0.37 296.25 1.62 0.29 301.25 1.23 0.25 309.75 1.77 0.20 314.75 1.36 0.23 319.75 1.85 0.32 324.75 1.87 0.39 329.75 3.84 0.69 330.75 6.00 0.93 334.75 2.08 0.38 339.75 1.92 0.31 341.75 2.04 0.36 344.75 2.21 0.39 345.75 2.55 0.37 354.75 2.36 0.38 359.75 2.08 0.45 364.75 2.26 0.46 369.75 2.05 0.40 374.75 2.04 0.63 377.75 1.74 0.22 379.75 1.93 0.55 382.75 2.01 0.42 384.75 1.67 0.37 389.75 1.48 0.33 394.75 1.72 0.37 397.75 1.24 0.27 399.75 0.99 0.28 404.75 1.15 0.21 407.75 1.75 0.48 409.75 1.18 0.21 414.75 1.27 0.27 419.75 0.72 0.11 424.75 0.58 0.14 429.75 0.90 0.28 434.75 0.91 0.28 436.75 0.65 0.27